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 46 毫秒
1.
2.
3.
4.
5.
Oligotrophic oceanic waters of the central ocean gyres typically have extremely low dissolved fixed inorganic nitrogen concentrations, but few nitrogen-fixing microorganisms from the oceanic environment have been cultivated. Nitrogenase gene (nifH) sequences amplified directly from oceanic waters showed that the open ocean contains more diverse diazotrophic microbial populations and more diverse habitats for nitrogen fixers than previously observed by classical microbiological techniques. Nitrogenase genes derived from unicellular and filamentous cyanobacteria, as well as from the α and γ subdivisions of the class Proteobacteria, were found in both the Atlantic and Pacific oceans. nifH sequences that cluster phylogenetically with sequences from sulfate reducers or clostridia were found associated with planktonic crustaceans. Nitrogenase sequence types obtained from invertebrates represented phylotypes distinct from the phylotypes detected in the picoplankton size fraction. The results indicate that there are in the oceanic environment several distinct potentially nitrogen-fixing microbial assemblages that include representatives of diverse phylotypes.The productivity of the oceans controls the fluxes of many biogeochemically important compounds, including the rate of exchange of carbon dioxide between the open ocean and the atmosphere. In turn, oceanic carbon fixation is limited by the bioavailability of nutrients, including nitrogen, phosphorus, and iron (9, 10, 20). In contrast to the biogeochemical cycles of phosphorus and iron, nitrogen is present in relatively high concentrations in seawater as gaseous N2. Gaseous nitrogen is available only to microorganisms with the capability of biological nitrogen fixation, the reduction of atmospheric N2 to ammonium. Although large areas of the world’s oceans are virtually devoid of fixed dissolved inorganic nitrogen and primary production may be nitrogen limited, very few species of nitrogen-fixing organisms have been identified or isolated from the plankton. Trichodesmium, a filamentous aggregate-forming cyanobacterium, is an abundant diazotroph in tropical and subtropical waters (3, 5), but few other examples of diazotrophs from the open ocean are known (21, 35). The seeming low diversity of known nitrogen-fixing organisms in the open ocean stands in stark contrast to the presumptive nitrogen limitation in the world’s oceans and presents an evolutionary paradox.Recently, biological nitrogen fixation has gained recognition as an important source of nitrogen for supporting oceanic primary production (3, 11, 18, 22). The nitrogen budget for the Atlantic Ocean does not balance because a source of nitrogen cannot be accounted for by current knowledge of fluxes and pools of nitrogen, even after including nitrogen fixation by Trichodesmium (22). It is speculated that rates of nitrogen fixation by known diazotrophic organisms have been underestimated (17), or as yet unidentified diazotrophic organisms are active in the ocean (18). Conventional nitrogenase, the enzyme that catalyzes biological dinitrogen reduction to ammonium, is composed of two highly conserved proteins: the iron (Fe) protein (encoded by the nifH gene) and the molybdenum iron (MoFe) protein (encoded by the nifDK genes). The nitrogenase enzyme is present in diverse lineages of prokaryotes and is generally believed to be ancient (38). Evolutionarily conserved amino acid sequences within the nifH (which encodes the Fe protein component of nitrogenase) gene have been exploited to design PCR primers to detect the genetic potential for nitrogen fixation in the marine environment (39). With this approach, the diversity of nitrogen-fixing microorganisms in oceanic water and marine plankton was determined. This report shows that there are far more diverse nitrogen-fixing populations and diverse habitats which can support nitrogen fixation in the open ocean than previously documented.  相似文献   

6.
7.
8.
9.
Entamoeba histolytica, the protist that causes amebic dysentery and liver abscess, has a truncated Asn-linked glycan (N-glycan) precursor composed of seven sugars (Man5GlcNAc2). Here, we show that glycoproteins with unmodified N-glycans are aggregated and capped on the surface of E. histolytica trophozoites by the antiretroviral lectin cyanovirin-N and then replenished from large intracellular pools. Cyanovirin-N cocaps the Gal/GalNAc adherence lectin, as well as glycoproteins containing O-phosphodiester-linked glycans recognized by an anti-proteophosphoglycan monoclonal antibody. Cyanovirin-N inhibits phagocytosis by E. histolytica trophozoites of mucin-coated beads, a surrogate assay for amebic virulence. For technical reasons, we used the plant lectin concanavalin A rather than cyanovirin-N to enrich secreted and membrane proteins for mass spectrometric identification. E. histolytica glycoproteins with occupied N-glycan sites include Gal/GalNAc lectins, proteases, and 17 previously hypothetical proteins. The latter glycoproteins, as well as 50 previously hypothetical proteins enriched by concanavalin A, may be vaccine targets as they are abundant and unique. In summary, the antiretroviral lectin cyanovirin-N binds to well-known and novel targets on the surface of E. histolytica that are rapidly replenished from large intracellular pools.Entamoeba histolytica causes amebic dysentery and liver abscess in the developing world (10, 20, 29). We are interested in E. histolytica glycoproteins containing Asn-linked glycans (N-glycans) for numerous reasons. E. histolytica makes an N-glycan precursor that contains 7 sugars (Man5GlcNAc2-PP-dolichol) rather than 14 sugars (Glc3Man9GlcNAc2-PP-dolichol) made by most animals, plants, and fungi (21, 31, 44). E. histolytica N-glycans are used for quality control of glycoprotein folding in the endoplasmic reticulum (ER) lumen, and there is positive selection for sites of N-linked glycosylation in secreted and membrane proteins of E. histolytica (5, 11, 53).Unprocessed Man5GlcNAc2, by far the most abundant E. histolytica N-glycan, is present on the plasma membrane and vesicular membranes (31). The antiretroviral lectin cyanovirin-N, which is specific for α-1,2-linked mannose present on unprocessed N-glycans, binds E. histolytica N-glycans and forms aggregates or caps on the surface of E. histolytica trophozoites (1, 25, 31, 44, 45). E. histolytica glycoproteins are also capped by the plant lectin concanavalin A (ConA), which has a broader carbohydrate specificity (mannose and glucose) than cyanovirin-N (3, 16, 18, 19). Heavy subunits of the Gal/GalNAc lectin, the most important E. histolytica vaccine candidate, have 7 to 10 potential sites for N-linked glycosylation (32, 39, 43). Inhibition of N-glycan synthesis results in Gal/GalNAc lectins that are unable to bind to sugars on host epithelial cells.Carbohydrates appear to be an important target on the surface of E. histolytica as anti-proteophosphoglycan (PPG) monoclonal antibodies bind to O-phosphodiester-linked glycans and protect animal models from amebic infection (6, 33, 35, 40, 48). Lectin affinity columns are a powerful method for enriching unique parasite glycoproteins that may be identified by mass spectrometry (MS) of tryptic fragments (17, 55). For example, we recently used the plant lectin wheat germ agglutinin to dramatically enrich glycoproteins with short N-glycans of Giardia (42).The goal of the present studies was to explore further the interaction of the antiretroviral lectin cyanovirin-N with E. histolytica trophozoites in vitro. Questions asked included the following: Are E. histolytica glycoproteins with N-glycans replenished on the plasma membrane after capping with cyanovirin-N? What is the effect of cyanovirin-N capping on other amebic virulence factors and/or vaccine candidates (e.g., the Gal/GalNAc lectin and PPG)? Is capping by cyanovirin-N mediated by actin, as described for capping by the Gal/GalNAc lectin and ConA? What is the effect of the cyanovirin-N on amebic phagocytosis of mucin-coated beads, a surrogate assay for virulence? Which trophozoite glycoproteins are potential targets of cyanovirin-N (identified by mass spectrometry of lectin-enriched E. histolytica proteins)? Are any of them potential vaccine candidates?  相似文献   

10.
In the endoplasmic reticulum (ER), lectins and processing enzymes are involved in quality control of newly synthesized proteins for productive folding as well as in the ER-associated degradation (ERAD) of misfolded proteins. ER quality control requires the recognition and modification of the N-linked oligosaccharides attached to glycoproteins. Mannose trimming from the N-glycans plays an important role in targeting of misfolded glycoproteins for ERAD. Recently, two mammalian lectins, OS-9 and XTP3-B, which contain mannose 6-phosphate receptor homology domains, were reported to be involved in ER quality control. Here, we examined the requirement for human OS-9 (hOS-9) lectin activity in degradation of the glycosylated ERAD substrate NHK, a genetic variant of α1-antitrypsin. Using frontal affinity chromatography, we demonstrated that the recombinant hOS-9 mannose 6-phosphate receptor homology domain specifically binds N-glycans lacking the terminal mannose from the C branch in vitro. To examine the specificity of OS-9 recognition of N-glycans in vivo, we modified the oligosaccharide structures on NHK by overexpressing ER α1,2-mannosidase I or EDEM3 and examined the effect of these modifications on NHK degradation in combination with small interfering RNA-mediated knockdown of hOS-9. The ability of hOS-9 to enhance glycoprotein ERAD depended on the N-glycan structures on NHK, consistent with the frontal affinity chromatography results. Thus, we propose a model for mannose trimming and the requirement for hOS-9 lectin activity in glycoprotein ERAD in which N-glycans lacking the terminal mannose from the C branch are recognized by hOS-9 and targeted for degradation.Recognition and sorting of improperly folded proteins is essential to cell survival, and hence, an elaborate quality control system is found in cells. ER4 quality control is well characterized with respect to the N-linked oligosaccharides regulating the folding and degradation of newly synthesized proteins in the ER (1). Immediately after polypeptides enter the ER, Glc3Man9GlcNAc2 (G3M9) precursor oligosaccharides are covalently attached and subsequently processed. Terminally misfolded proteins are removed from the ER by the ERAD machinery (14). Aberrant conformers are recognized, retrotranslocated to the cytosol, and degraded by the ubiquitin-proteasome system (5, 6). Processing of mannose residues from the N-linked oligosaccharides acts as a timer for the recognition of misfolded glycoproteins in the ER lumen (1, 7). ER α1,2-mannosidase I (ER ManI) in mammals and ER α-mannosidase in yeast preferentially trim mannose residues from the middle branch of N-glycans, generating the Man8GlcNAc2 (M8) isomer B (M8B) (8). In mammals, further mannose processing is required as a signal for degradation (1, 9, 10), whereas the presence of M8B is sufficient to signal degradation in yeast (11). The postulated lectin EDEMs in mammals, their yeast homolog Htm1p/Mnl1p, and the yeast MRH domain-containing lectin Yos9p have all been proposed to recognize glycoproteins targeted for degradation (12).The role of Yos9p in glycoprotein ERAD was identified using a genetic screen in Saccharomyces cerevisiae (13). Yos9p, a homolog of hOS-9, contains an MRH domain (14) and functions as a lectin. Yos9p recognizes substrates of the ERAD-lumenal pathway (1517), generating a large ER membrane complex containing the Hrd1p-Hrd3p ubiquitin ligase core complex (1820). The M8B and Man5GlcNAc2 (M5) N-glycans are predicted to function as ligands for Yos9p (17). Bipartite recognition of both glycan and polypeptide by Yos9p has also been reported (15).Recent studies revealed that two mammalian MRH domain-containing lectins, OS-9 and XTP3-B, are ER luminal proteins involved in ER quality control and form a large complex containing the HRD1-SEL1L ubiquitin-ligase in the ER membrane (2124). The components of the complex are similar to yeast, suggesting evolutionary conservation, although the molecular mechanisms underlying the role of OS-9 and XTP3-B remain elusive. Studies using lectin mutants have suggested that the MRH domains are required not for binding to ERAD substrates but for interactions with SEL1L (21), which has multiple N-glycans (25, 26). Additionally, lectin activity appears to be dispensable for hOS-9 binding to misfolded glycoproteins (21, 24). Thus, to understand the role of hOS-9 in the ER quality control pathway, the specific carbohydrate structures recognized by the hOS-9 MRH domain need to be identified, and the requirement of the lectin domain in substrate recognition needs to be determined.In the present study we demonstrate that the lectin activity of hOS-9 is required for enhancement of glycoprotein ERAD. We identified the N-glycan structures recognized by the recombinant hOS-9 MRH domain in vitro by frontal affinity chromatography (FAC). Using a model ERAD substrate, NHK (27), we show that the ability of hOS-9 to enhance ERAD in vivo depends on the oligosaccharides present on NHK, consistent with the FAC results.  相似文献   

11.
This paper presents evidence that a member of the L1 family of ankyrin-binding cell adhesion molecules is a substrate for protein tyrosine kinase(s) and phosphatase(s), identifies the highly conserved FIGQY tyrosine in the cytoplasmic domain as the principal site of phosphorylation, and demonstrates that phosphorylation of the FIGQY tyrosine abolishes ankyrin-binding activity. Neurofascin expressed in neuroblastoma cells is subject to tyrosine phosphorylation after activation of tyrosine kinases by NGF or bFGF or inactivation of tyrosine phosphatases with vanadate or dephostatin. Furthermore, both neurofascin and the related molecule Nr-CAM are tyrosine phosphorylated in a developmentally regulated pattern in rat brain. The FIGQY sequence is present in the cytoplasmic domains of all members of the L1 family of neural cell adhesion molecules. Phosphorylation of the FIGQY tyrosine abolishes ankyrin binding, as determined by coimmunoprecipitation of endogenous ankyrin and in vitro ankyrin-binding assays. Measurements of fluorescence recovery after photobleaching demonstrate that phosphorylation of the FIGQY tyrosine also increases lateral mobility of neurofascin expressed in neuroblastoma cells to the same extent as removal of the cytoplasmic domain. Ankyrin binding, therefore, appears to regulate the dynamic behavior of neurofascin and is the target for regulation by tyrosine phosphorylation in response to external signals. These findings suggest that tyrosine phosphorylation at the FIGQY site represents a highly conserved mechanism, used by the entire class of L1-related cell adhesion molecules, for regulation of ankyrin-dependent connections to the spectrin skeleton.Vertebrate L1, neurofascin, neuroglial cell adhesion molecule (Ng-CAM),1 Ng-CAM–related cell adhesion molecule (Nr-CAM), and Drosophila neuroglian are members of a family of nervous system cell adhesion molecules that possess variable extracellular domains comprised of Ig and fibronectin type III domains and a relatively conserved cytoplasmic domain (Grumet, 1991; Hortsch and Goodman, 1991; Rathgen and Jessel, 1991; Sonderegger and Rathgen, 1992; Hortsch, 1996). Members of this family, including a number of alternatively spliced forms, are abundant in the nervous system during early development as well as in adults. Neurofascin and Nr-CAM, for example, constitute ∼0.5% of the total membrane protein in adult brain (Davis et al., 1993; Davis and Bennett, 1994). Cellular functions attributed to the L1 family include axon fasciculation (Stallcup and Beasley, 1985; Landmesser et al., 1988; Brummendorf and Rathjen, 1993; Bastmeyer et al., 1995; Itoh et al., 1995; Magyar-Lehmann et al., 1995), axonal guidance (van den Pol and Kim, 1993; Liljelund et al., 1994; Brittis and Silver, 1995; Brittis et al., 1995; Lochter et al., 1995; Wong et al., 1996), neurite extension (Chang et al., 1987; Felsenfeld et al., 1994; Hankin and Lagenaur, 1994; Ignelzi et al., 1994; Williams et al., 1994a ,b,c,d; Doherty et al., 1995; Zhao and Siu, 1995), a role in long term potentiation (Luthl et al., 1994), synaptogenesis (Itoh et al., 1995), and myelination (Wood et al., 1990). The potential clinical importance of this group of proteins has been emphasized by the findings that mutations in the L1 gene on the X chromosome are responsible for developmental anomalies including hydrocephalus and mental retardation (Rosenthal et al., 1992; Jouet et al., 1994; Wong et al., 1995).The conserved cytoplasmic domains of L1 family members include a binding site for the membrane skeletal protein ankyrin. This interaction was first described for neurofascin (Davis et. al., 1993) and subsequently has been observed for L1, Nr-CAM (Davis and Bennett, 1994), and Drosophila neuroglian (Dubreuil et al., 1996). The membrane-binding domain of ankyrin contains two distinct sites for neurofascin and has the potential to promote lateral association of neurofascin and presumably other L1 family members (Michaely and Bennett, 1995). Nodes of Ranvier are physiologically relevant axonal sites where ankyrin and L1 family members collaborate, based on findings of colocalization of a specialized isoform of ankyrin with alternatively spliced forms of neurofascin and NrCAM in adults (Davis et al., 1996) as well as in early axonal developmental intermediates (Lambert, S., J. Davis, P. Michael, and V. Bennett. 1995. Mol. Biol. Cell. 6:98a).L1, after homophilic and/or heterophilic binding, participates in signal transduction pathways that ultimately are associated with neurite extension and outgrowth (Ignelzi et al., 1994; Williams et al., 1994a ,b,c,d; Doherty et al., 1995). L1 copurifies with a serine–threonine protein kinase (Sadoul et al., 1989) and is phosphorylated on a serine residue that is not conserved among other family members (Wong et al., 1996). L1 pathway(s) may also involve G proteins, calcium channels, and tyrosine phosphorylation (Williams et al., 1994a ,b,c,d; Doherty et al., 1995). After homophilic interactions, L1 directly activates a tyrosine signaling cascade after a lateral association of its ectodomain with the fibroblast growth factor receptor (Doherty et al., 1995). Antibodies against L1 have also been shown to activate protein tyrosine phosphatase activity in growth cones (Klinz et al., 1995). However, details of the downstream substrates of L1-promoted phosphorylation and dephosphorylation and possible roles of the cytoplasmic domain are not known.Tyrosine phosphorylation is well established to modulate cell–cell and cell–extracellular matrix interactions involving integrins and their associated proteins (Akiyama et al., 1994; Arroyo et al., 1994; Schlaepfer et al., 1994; Law et al., 1996) as well as the cadherins (Balsamo et al., 1996; Krypta et al., 1996; Brady-Kalnay et al., 1995; Shibamoto et al., 1995; Hoschuetzky et al., 1994; Matsuyoshi et al., 1992). For example, the adhesive functions of the calciumdependent cadherin cell adhesion molecule are mediated by a dynamic balance between tyrosine phosphorylation of β-catenin by TrkA and dephosphorylation via the LARtype protein tyrosine phosphatase (Krypta et al., 1996). In this example the regulation of binding among the structural proteins is the result of a coordination between classes of protein kinases and protein phosphatases.This study presents evidence that neurofascin, expressed in a rat neuroblastoma cell line, is a substrate for both tyrosine kinases and protein tyrosine phosphatases at a tyrosine residue conserved among all members of the L1 family. Site-specific tyrosine phosphorylation promoted by both tyrosine kinase activators (NGF and bFGF) and protein tyrosine phosphatase inhibitors (dephostatin and vanadate) is a strong negative regulator of the neurofascin– ankyrin binding interaction and modulates the membrane dynamic behavior of neurofascin. Furthermore, neurofascin and, to a lesser extent Nr-CAM, are also shown here to be tyrosine phosphorylated in developing rat brain, implying a physiological relevance to this phenomenon. These results indicate that neurofascin may be a target for the coordinate control over phosphorylation that is elicited by protein kinases and phosphatases during in vivo tyrosine phosphorylation cascades. The consequent decrease in ankyrin-binding capacity due to phosphorylation of neurofascin could represent a general mechanism among the L1 family members for regulation of membrane–cytoskeletal interactions in both developing and adult nervous systems.  相似文献   

12.
Isochorismate is an important metabolite formed at the end of the shikimate pathway, which is involved in the synthesis of both primary and secondary metabolites. It is synthesized from chorismate in a reaction catalyzed by the enzyme isochorismate synthase (ICS; EC 5.4.99.6). We have purified ICS to homogeneity from elicited Catharanthus roseus cell cultures. Two isoforms with an apparent molecular mass of 64 kD were purified and characterized. The Km values for chorismate were 558 and 319 μm for isoforms I and II, respectively. The isoforms were not inhibited by aromatic amino acids and required Mg2+ for enzyme activity. Polymerase chain reaction on a cDNA library from elicited C. roseus cells with a degenerated primer based on the sequence of an internal peptide from isoform II resulted in an amplification product that was used to screen the cDNA library. This led to the first isolation, to our knowledge, of a plant ICS cDNA. The cDNA encodes a protein of 64 kD with an N-terminal chloroplast-targeting signal. The deduced amino acid sequence shares homology with bacterial ICS and also with anthranilate synthases from plants. Southern analysis indicates the existence of only one ICS gene in C. roseus.The shikimate pathway is a major pathway in primary and secondary plant metabolism (Herrmann, 1995). It provides chorismate for the synthesis of the aromatic amino acids Phe, Tyr, and Trp, which are used in protein biosynthesis, but also serves as a precursor for a wide variety of aromatic substances (Herrmann, 1995; Weaver and Hermann, 1997; Fig. Fig.1a).1a). Chorismate is also the starting point of a biosynthetic pathway leading to phylloquinones (vitamin K1) and anthraquinones (Poulsen and Verpoorte, 1991). The first committed step in this pathway is the conversion of chorismate into isochorismate, which is catalyzed by ICS (Poulsen and Verpoorte, 1991; Fig. Fig.1b).1b). Its substrate, chorismate, plays a pivotal role in the synthesis of shikimate-pathway-derived compounds, and its distribution over the various pathways is expected to be tightly regulated. Elicited cell cultures of Catharanthus roseus provide an example of the partitioning of chorismate. Concurrently, these cultures produce both Trp-derived indole alkaloids and DHBA (Moreno et al., 1994). In bacteria DHBA is synthesized from isochorismate (Young et al., 1969). Elicitation of C. roseus cell cultures with a fungal extract induces not only several enzymes of the indole alkaloid biosynthetic pathway (Pasquali et al., 1992) but also ICS (Moreno et al., 1994). Information concerning the expression and biochemical characteristics of the enzymes that compete for available chorismate (ICS, CM, and AS) may help us to understand the regulation of the distribution of this precursor over the various pathways. Such information is already available for CM (Eberhard et al., 1996) and AS (Poulsen et al., 1993; Bohlmann et al., 1995) but not for ICS. Figure 1a, Position of ICS in the plant metabolism. SA, Salicylic acid, OSB, o-succinylbenzoic acid. b, Reaction catalyzed by ICS.Isochorismate plays an important role in bacterial and plant metabolism as a precursor of o-succinylbenzoic acid, an intermediate in the biosynthesis of menaquinones (vitamin K2) (Weische and Leistner, 1985) and phylloquinones (vitamin K1; Poulsen and Verpoorte, 1991). In bacteria isochorismate is also a precursor of siderophores such as DHBA (Young et al., 1969), enterobactin (Walsh et al., 1990), amonabactin (Barghouthi et al., 1991), and salicylic acid (Serino et al., 1995). Although evidence from tobacco would indicate that salicylic acid in plants is derived from Phe via benzoic acid (Yalpani et al., 1993; Lee et al., 1995; Coquoz et al., 1998), it cannot be excluded that it is also synthesized from isochorismate. In the secondary metabolism of higher plants, isochorismate is a precursor for the biosynthesis of anthraquinones (Inoue et al., 1984; Sieweke and Leistner, 1992), naphthoquinones (Müller and Leistner, 1978), catalpalactone (Inouye et al., 1975), and certain alkaloids in orchids (Leete and Bodem, 1976).ICS was first extracted and partially purified from crude extracts of Aerobacter aerogenes (Young and Gibson, 1969). Later, ICS activity was detected in protein extracts of cell cultures from plants of the Rubiaceae, Celastraceae, and Apocynaceae families (Ledüc et al., 1991; Poulsen et al., 1991; Poulsen and Verpoorte, 1992). Genes encoding ICS have been cloned from bacteria such as Escherichia coli (Ozenberger et al., 1989), Pseudomonas aeruginosa (Serino et al., 1995), Aeromonas hydrophila (Barghouthi et al., 1991), Flavobacterium K3–15 (Schaaf et al., 1993), Hemophilus influenzae (Fleischmann et al., 1995), and Bacillus subtilis (Rowland and Taber, 1996). Both E. coli and B. subtilis have two distinct ICS genes; one is involved in siderophore biosynthesis and the other is involved in menaquinone production (Daruwala et al., 1996, 1997; Müller et al., 1996; Rowland and Taber, 1996). The biochemical properties of the two ICS enzymes from E. coli are different (Daruwala et al., 1997; Liu et al., 1990). Sequence analysis has revealed that the bacterial ICS enzymes share homology with the chorismate-utilizing enzymes AS and p-aminobenzoate synthase, suggesting that they share a common evolutionary origin (Ozenberger et al., 1989).Much biochemical and molecular data concerning the shikimate pathway in plants have accumulated in recent years (Schmid and Amrhein, 1995; Weaver and Hermann, 1997), but relatively little work has been done on ICS from higher plants. The enzyme has been partially purified from Galium mollugo cell cultures (Ledüc et al., 1991, 1997), but purification of the ICS protein to homogeneity has remained elusive, probably because of instability of the enzyme.Our interests focus on the role of ICS in the regulation of chorismate partitioning over the various pathways. Furthermore, we studied ICS in C. roseus to gain insight into the biosynthesis of DHBA in higher plants (Moreno et al., 1994). In this paper we report the first purification, to our knowledge, of ICS to homogeneity from a plant source and the cloning of the corresponding cDNA.  相似文献   

13.
14.
15.
Glycoprotein structure determination and quantification by MS requires efficient isolation of glycopeptides from a proteolytic digest of complex protein mixtures. Here we describe that the use of acids as ion-pairing reagents in normal-phase chromatography (IP-NPLC) considerably increases the hydrophobicity differences between non-glycopeptides and glycopeptides, thereby resulting in the reproducible isolation of N-linked high mannose type and sialylated glycopeptides from the tryptic digest of a ribonuclease B and fetuin mixture. The elution order of non-glycopeptides relative to glycopeptides in IP-NPLC is predictable by their hydrophobicity values calculated using the Wimley-White water/octanol hydrophobicity scale. O-linked glycopeptides can be efficiently isolated from fetuin tryptic digests using IP-NPLC when N-glycans are first removed with PNGase. IP-NPLC recovers close to 100% of bacterial N-linked glycopeptides modified with non-sialylated heptasaccharides from tryptic digests of periplasmic protein extracts from Campylobacter jejuni 11168 and its pglD mutant. Label-free nano-flow reversed-phase LC-MS is used for quantification of differentially expressed glycopeptides from the C. jejuni wild-type and pglD mutant followed by identification of these glycoproteins using multiple stage tandem MS. This method further confirms the acetyltransferase activity of PglD and demonstrates for the first time that heptasaccharides containing monoacetylated bacillosamine are transferred to proteins in both the wild-type and mutant strains. We believe that IP-NPLC will be a useful tool for quantitative glycoproteomics.Protein glycosylation is a biologically significant and complex post-translational modification, involved in cell-cell and receptor-ligand interactions (14). In fact, clinical biomarkers and therapeutic targets are often glycoproteins (59). Comprehensive glycoprotein characterization, involving glycosylation site identification, glycan structure determination, site occupancy, and glycan isoform distribution, is a technical challenge particularly for quantitative profiling of complex protein mixtures (1013). Both N- and O-glycans are structurally heterogeneous (i.e. a single site may have different glycans attached or be only partially occupied). Therefore, the MS1 signals from glycopeptides originating from a glycoprotein are often weaker than from non-glycopeptides. In addition, the ionization efficiency of glycopeptides is low compared with that of non-glycopeptides and is often suppressed in the presence of non-glycopeptides (1113). When the MS signals of glycopeptides are relatively high in simple protein digests then diagnostic sugar oxonium ion fragments produced by, for example, front-end collisional activation can be used to detect them. However, when peptides and glycopeptides co-elute, parent ion scanning is required to selectively detect the glycopeptides (14). This can be problematic in terms of sensitivity, especially for detecting glycopeptides in digests of complex protein extracts.Isolation of glycopeptides from proteolytic digests of complex protein mixtures can greatly enhance the MS signals of glycopeptides using reversed-phase LC-ESI-MS (RPLC-ESI-MS) or MALDI-MS (1524). Hydrazide chemistry is used to isolate, identify, and quantify N-linked glycopeptides effectively, but this method involves lengthy chemical procedures and does not preserve the glycan moieties thereby losing valuable information on glycan structure and site occupancy (1517). Capturing glycopeptides with lectins has been widely used, but restricted specificities and unspecific binding are major drawbacks of this method (1821). Under reversed-phase LC conditions, glycopeptides from tryptic digests of gel-separated glycoproteins have been enriched using graphite powder medium (22). In this case, however, a second digestion with proteinase K is required for trimming down the peptide moieties of tryptic glycopeptides so that the glycopeptides (typically <5 amino acid residues) essentially resemble the glycans with respect to hydrophilicity for subsequent separation. Moreover, the short peptide sequences of the proteinase K digest are often inadequate for de novo sequencing of the glycopeptides.Glycopeptide enrichment under normal-phase LC (NPLC) conditions has been demonstrated using various hydrophilic media and different capture and elution conditions (2328). NPLC allows either direct enrichment of peptides modified by various N-linked glycan structures using a ZIC®-HILIC column (2327) or targeting sialylated glycopeptides using a titanium dioxide micro-column (28). However, NPLC is neither effective for enriching less hydrophilic glycopeptides, e.g. the five high mannose type glycopeptides modified by 7–11 monosaccharide units from a tryptic digest of ribonuclease b (RNase B), nor for enriching O-linked glycopeptides of bovine fetuin using a ZIC-HILIC column (23). The use of Sepharose medium for enriching glycopeptides yielded only modest recovery of glycopeptides (28). In addition, binding of hydrophilic non-glycopeptides with these hydrophilic media contaminates the enriched glycopeptides (23, 28).We have recently developed an ion-pairing normal-phase LC (IP-NPLC) method to enrich glycopeptides from complex tryptic digests using Sepharose medium and salts or bases as ion-pairing reagents (29). Though reasonably effective the technique still left room for significant improvement. For example, the method demonstrated relatively modest glycopeptide selectivity, providing only 16% recovery for high mannose type glycopeptides (29). Here we report on a new IP-NPLC method using acids as ion-pairing reagents and polyhydroxyethyl aspartamide (A) as the stationary phase for the effective isolation of tryptic glycopeptides. The method was developed and evaluated using a tryptic digest of RNase B and fetuin mixture. In addition, we demonstrate that O-linked glycopeptides can be effectively isolated from a fetuin tryptic digest by IP-NPLC after removal of the N-linked glycans by PNGase F.The new IP-NPLC method was used to enrich N-linked glycopeptides from the tryptic digests of protein extracts of wild-type (wt) and PglD mutant strains of Campylobacter jejuni NCTC 11168. C. jejuni has a unique N-glycosylation system that glycosylates periplasmic and inner membrane proteins containing the extended N-linked sequon, D/E-X-N-X-S/T, where X is any amino acid other than proline (3032). The N-linked glycan of C. jejuni has been previously determined to be GalNAc-α1,4-GalNAc-α1,4-[Glcβ1,3]-GalNAc-α1,4-GalNAc-α1,4-GalNAc-α1,3-Bac-β1 (BacGalNAc5Glc residue mass: 1406 Da), where Bac is 2,4-diacetamido-2,4,6-trideoxyglucopyranose (30). In addition, the glycan structure of C. jejuni is conserved, unlike in eukaryotic systems (3032). IP-NPLC recovered close to 100% of the bacterial N-linked glycopeptides with virtually no contamination of non-glycopeptides. Furthermore, we demonstrate for the first time that acetylation of bacillosamine is incomplete in the wt using IP-NPLC and label-free MS.  相似文献   

16.
17.
18.
19.
20.
NAD-isocitrate dehydrogenase (NAD-IDH) from the eukaryotic microalga Chlamydomonas reinhardtii was purified to electrophoretic homogeneity by successive chromatography steps on Phenyl-Sepharose, Blue-Sepharose, diethylaminoethyl-Sephacel, and Sephacryl S-300 (all Pharmacia Biotech). The 320-kD enzyme was found to be an octamer composed of 45-kD subunits. The presence of isocitrate plus Mn2+ protected the enzyme against thermal inactivation or inhibition by specific reagents for arginine or lysine. NADH was a competitive inhibitor (Ki, 0.14 mm) and NADPH was a noncompetitive inhibitor (Ki, 0.42 mm) with respect to NAD+. Citrate and adenine nucleotides at concentrations less than 1 mm had no effect on the activity, but 10 mm citrate, ATP, or ADP had an inhibitory effect. In addition, NAD-IDH was inhibited by inorganic monovalent anions, but l-amino acids and intermediates of glycolysis and the tricarboxylic acid cycle had no significant effect. These data support the idea that NAD-IDH from photosynthetic organisms may be a key regulatory enzyme within the tricarboxylic acid cycle.IDH catalyzes the oxidative decarboxylation of isocitrate to produce 2-oxoglutarate. According to the specificity for the electron acceptor, two enzymes with IDH activity are known, NAD-IDH (EC 1.1.1.41) and NADP-IDH (EC 1.1.1.42) (Chen and Gadal, 1990a).In photosynthetic organisms NADP-IDH has been detected in the cytosol, chloroplasts, mitochondria, and peroxisomes. Cytosolic NADP-IDH has been purified from higher plants (Chen et al., 1988) and eukaryotic algae (Martínez-Rivas et al., 1996), and its cDNA has been cloned from alfalfa (Shorrosh and Dixon, 1992), soybean (Udvardi et al., 1993), potato (Fieuw et al., 1995), and tobacco (Gálvez et al., 1996). This 80-kD isoenzyme is a dimer, and it is likely to be involved in the synthesis of NADPH for biosynthetic purposes in the cytosol (Chen et al., 1988), in the synthesis of 2-oxoglutarate for ammonium assimilation (Chen and Gadal, 1990b), and in the cycling, redistribution, and export of amino acids (Fieuw et al., 1995). Chloroplastic NADP-IDH has been studied in higher plants (Gálvez et al., 1994) and eukaryotic algae (Martínez-Rivas and Vega, 1994). It is a 154-kD dimer that has been proposed to be involved in the supply of NADPH for biosynthetic reactions in the chloroplast when photosynthetic NADPH production is low (Gálvez et al., 1994). The mitochondrial NADP-IDH of higher plants may have a physiological role in the production of NADPH, which can be converted to NADH by a transhydrogenase or used to reduce glutathione in the mitochondrial matrix (Rasmusson and Møller, 1990). NADP-IDH activity has also been detected in peroxisomes from spinach leaves (Yamazaki and Tolbert, 1970).NAD-IDH is localized exclusively in the mitochondria in association with the TCA cycle. This enzyme has been purified from several nonphotosynthetic eukaryotes such as fungi (Keys and McAlister-Henn, 1990; Alvarez-Villafañe et al., 1996) and animals (Giorgio et al., 1970), in which it appears to be a 300-kD octamer. Its key regulatory role in the TCA cycle is well documented. The NAD-IDH from yeast is activated by AMP and citrate (Hathaway and Atkinson, 1963), whereas the animal enzyme is activated by ADP and citrate (Cohen and Colman, 1972). In addition, the NAD-IDH cDNAs have been cloned from yeast (Cupp and McAlister-Henn, 1991, 1992) and animals (Nichols et al., 1995; Zeng et al., 1995). In these organisms, the enzyme is composed of two (yeast) or more (animals) different subunits encoded by different genes.To our knowledge, no NAD-IDH from photosynthetic organisms has yet been purified to homogeneity, mainly because of the low stability of the enzyme (Oliver and McIntosh, 1995). However, partial purifications have been reported from pea (Cox and Davies, 1967; Cox, 1969; McIntosh and Oliver, 1992), potato (Laties, 1983), spruce (Cornu et al., 1996), and the eukaryotic microalga Chlamydomonas reinhardtii (Martínez-Rivas and Vega, 1994). Matrix and membrane forms of the enzyme have been detected in potato (Tezuka and Laties, 1983) and pea (McIntosh, 1997). Although it is an allosteric enzyme that exhibits sigmoidal kinetics with respect to isocitrate (Cox and Davies, 1967; McIntosh and Oliver, 1992) and is activated in vitro by ABA (Tezuka et al., 1990), the regulatory importance of NAD-IDH in photosynthetic organisms is still under debate.To elucidate the regulatory significance of NAD-IDH in photosynthetic organisms and its apparent contribution to the 2-oxoglutarate supply for ammonium assimilation, we have purified and characterized the NAD-IDH from C. reinhardtii.  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