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 406 毫秒
1.
2.
This study demonstrates the utility of Lifeact for the investigation of actin dynamics in Neurospora crassa and also represents the first report of simultaneous live-cell imaging of the actin and microtubule cytoskeletons in filamentous fungi. Lifeact is a 17-amino-acid peptide derived from the nonessential Saccharomyces cerevisiae actin-binding protein Abp140p. Fused to green fluorescent protein (GFP) or red fluorescent protein (TagRFP), Lifeact allowed live-cell imaging of actin patches, cables, and rings in N. crassa without interfering with cellular functions. Actin cables and patches localized to sites of active growth during the establishment and maintenance of cell polarity in germ tubes and conidial anastomosis tubes (CATs). Recurrent phases of formation and retrograde movement of complex arrays of actin cables were observed at growing tips of germ tubes and CATs. Two populations of actin patches exhibiting slow and fast movement were distinguished, and rapid (1.2 μm/s) saltatory transport of patches along cables was observed. Actin cables accumulated and subsequently condensed into actin rings associated with septum formation. F-actin organization was markedly different in the tip regions of mature hyphae and in germ tubes. Only mature hyphae displayed a subapical collar of actin patches and a concentration of F-actin within the core of the Spitzenkörper. Coexpression of Lifeact-TagRFP and β-tubulin–GFP revealed distinct but interrelated localization patterns of F-actin and microtubules during the initiation and maintenance of tip growth.Actins are highly conserved proteins found in all eukaryotes and have an enormous variety of cellular roles. The monomeric form (globular actin, or G-actin) can self-assemble, with the aid of numerous actin-binding proteins (ABPs), into microfilaments (filamentous actin, or F-actin), which, together with microtubules, form the two major components of the fungal cytoskeleton. Numerous pharmacological and genetic studies of fungi have demonstrated crucial roles for F-actin in cell polarity, exocytosis, endocytosis, cytokinesis, and organelle movement (6, 7, 20, 34, 35, 51, 52, 59). Phalloidin staining, immunofluorescent labeling, and fluorescent-protein (FP)-based live-cell imaging have revealed three distinct subpopulations of F-actin-containing structures in fungi: patches, cables, and rings (1, 14, 28, 34, 60, 63, 64). Actin patches are associated with the plasma membrane and represent an accumulation of F-actin around endocytic vesicles (3, 26, 57). Actin cables are bundles of actin filaments stabilized with cross-linking proteins, such as tropomyosins and fimbrin, and are assembled by formins at sites of active growth, where they form tracks for myosin V-dependent polarized secretion and organelle transport (10, 16, 17, 27, 38, 47, 48). Cables, unlike patches, are absolutely required for polarized growth in the budding yeast Saccharomyces cerevisiae (34, 38). Contractile actomyosin rings are essential for cytokinesis in budding yeast, whereas in filamentous fungi, actin rings are less well studied but are known to be involved in septum formation (20, 28, 34, 39, 40).Actin cables and patches have been particularly well studied in budding yeast. However, there are likely to be important differences between F-actin architecture and dynamics in budding yeast and those in filamentous fungi, as budding yeasts display only a short period of polarized growth during bud formation, which is followed by isotropic growth over the bud surface (10). Sustained polarized growth during hyphal morphogenesis is a defining feature of filamentous fungi (21), making them attractive models for studying the roles of the actin cytoskeleton in cell polarization, tip growth, and organelle transport.In Neurospora crassa and other filamentous fungi, disruption of the actin cytoskeleton leads to rapid tip swelling, which indicates perturbation of polarized tip growth, demonstrating a critical role for F-actin in targeted secretion to particular sites on the plasma membrane (7, 22, 29, 56). Immunofluorescence studies of N. crassa have shown that F-actin localizes to hyphal tips as “clouds” and “plaques” (7, 54, 59). However, immunolabeling has failed to reveal actin cables in N. crassa and offers limited insights into F-actin dynamics. Live-cell imaging of F-actin architecture and dynamics has not been accomplished in N. crassa, yet it is expected to yield key insights into cell polarization, tip growth, and intracellular transport.We took advantage of a recently developed live-cell imaging probe for F-actin called Lifeact (43). Lifeact is a 17-amino-acid peptide derived from the N terminus of the budding yeast actin-binding protein Abp140 (5, 63) and has recently been demonstrated to be a universal live-cell imaging marker for F-actin in eukaryotes (43). Here, we report the successful application of fluorescent Lifeact fusion constructs for live-cell imaging of F-actin in N. crassa. We constructed two synthetic genes consisting of Lifeact fused to “synthetic” green fluorescent protein (sGFP) (S65T) (henceforth termed GFP) (12) or red fluorescent protein (TagRFP) (33) and expressed these constructs in various N. crassa strains. In all strain backgrounds, fluorescent Lifeact constructs clearly labeled actin patches, cables, and rings and revealed a direct association of F-actin structures with sites of cell polarization and active tip growth. Our results demonstrate the efficacy of Lifeact as a nontoxic live-cell imaging probe in N. crassa.  相似文献   

3.
Fusarium oxysporum is the causative agent of fungal wilt disease in a variety of crops. The capacity of a fungal pathogen such as F. oxysporum f. sp. nicotianae to establish infection on its tobacco (Nicotiana tabacum) host depends in part on its capacity to evade the toxicity of tobacco defense proteins, such as osmotin. Fusarium genes that control resistance to osmotin would therefore reflect coevolutionary pressures and include genes that control mutual recognition, avoidance, and detoxification. We identified FOR (Fusarium Osmotin Resistance) genes on the basis of their ability to confer osmotin resistance to an osmotin-sensitive strain of Saccharomyces cerevisiae. FOR1 encodes a putative cell wall glycoprotein. FOR2 encodes the structural gene for glutamine:fructose-6-phosphate amidotransferase, the first and rate-limiting step in the biosynthesis of hexosamine and cell wall chitin. FOR3 encodes a homolog of SSD1, which controls cell wall composition, longevity, and virulence in S. cerevisiae. A for3 null mutation increased osmotin sensitivity of conidia and hyphae of F. oxysporum f. sp. nicotianae and also reduced cell wall β-1,3-glucan content. Together our findings show that conserved fungal genes that determine cell wall properties play a crucial role in regulating fungal susceptibility to the plant defense protein osmotin.Studies of plant-pathogen interactions strongly suggest that under the pressure to survive, plants and pathogens continuously react to one another''s defense arsenal and evolve to overcome these defenses (13). Plants recognize pathogen-associated molecular patterns, such as fungal cell wall fragments composed of chitin, glucans, oligosaccharides, or glycoprotein peptides (32). It has been established that pathogens evolved effector proteins to avoid plant surveillance mechanisms that recognize pathogen-associated molecular patterns and this in turn led to the evolution of plant surveillance mechanisms that recognize pathogen-specific effector proteins. All pathogen recognition mechanisms induce intracellular signaling that culminates in the synthesis of factors, such as antimicrobial plant proteins, that help in limiting the severity of infection (74). The antimicrobial proteins are therefore among the ultimate effectors of plant defense. There is evidence of recognition between plant antimicrobial proteins and pathogen-specific molecules (74). Therefore, pathogen mechanisms of resistance to the antimicrobial proteins and the antimicrobial proteins themselves must have coevolved. Consequently, we postulated that a screen for fungal genes that alter the sensitivity of a phytopathogen to an antifungal protein of the host plant (that is, a cognate plant defense effector) would lead to identification of genes involved in controlling pathogenicity, in controlling access of the antifungal protein to its target fungal molecules (such as genes controlling cell surface composition), and in controlling detoxification mechanisms.The plant antifungal protein selected to test this hypothesis was osmotin. Osmotin is an antifungal protein that is overexpressed in and secreted by salt-adapted cultured tobacco (Nicotiana tabacum) cells (63). It is a member of a family of ubiquitous plant proteins, referred to as plant pathogenesis-related proteins of family 5 (PR-5), that are implicated in defense against fungi (74). Osmotin gene and protein expression is induced by biotic stresses, and overexpression of osmotin delays development of disease symptoms in transgenic plants (41, 42, 43, 84). The genetic bases of the susceptibility and resistance of Saccharomyces cerevisiae to osmotin have been explored in our laboratory (49, 50). The results show that specific interactions of osmotin with the plasma membrane are responsible for cell death signaling. However, because the cell wall governs access of osmotin to the plasma membrane, differences in cell wall composition largely account for the differential osmotin sensitivity of various S. cerevisiae strains, and specific cell wall components play a significant role in modulating osmotin toxicity (30, 31, 49, 50, 81, 82). These studies in the model nonpathogenic fungus, S. cerevisiae, support our hypothesis that a screen for genes that alter the sensitivity of a phytopathogenic fungus to an antifungal defense effector protein of the host plant will uncover genes involved in controlling access of the antifungal protein to its target fungal molecules.Osmotin, like other plant defense antifungal proteins, has specific but broad-spectrum antifungal activity (74). One of the most osmotin-sensitive phytopathogenic fungi is Fusarium oxysporum. F. oxysporum is an ascomycete fungus, like S. cerevisiae, and has been touted as an appropriate multihost model for studying fungal virulence (53). It is a soilborne plant pathogen of economic significance, because it causes vascular wilt disease on a large variety of crop plants and produces toxic food contaminants (17, 58). In humans it also causes skin, nail, and eye disease that can become serious or life-threatening illnesses in immunocompromised patients (52, 69). F. oxysporum f. sp. lycopersici, F. oxysporum f. sp. nicotianae, and F. oxysporum f. sp. meloni, like S. cerevisiae, are quite sensitive to osmotin (1, 51; M. L. Narasimhan, unpublished data). Furthermore, it was recently shown that overexpression in F. oxysporum f. sp. nicotianae of an S. cerevisiae cell wall glycoprotein that increases the osmotin resistance of S. cerevisiae also increases the osmotin resistance of the plant pathogen and its virulence on tobacco, the osmotin-producing host plant (51). This suggested that osmotin resistance mechanisms may be conserved between S. cerevisiae and F. oxysporum and that S. cerevisiae could be used as a tool to uncover F. oxysporum genes that control osmotin sensitivity or resistance.In the current study, we expressed an F. oxysporum f. sp. nicotianae cDNA library in the osmotin-sensitive S. cerevisiae strain BWG1-7a and selected genes for their ability to increase osmotin tolerance. We report here the identification and characterization of three FOR (Fusarium Osmotin Resistance) genes that affect the cell wall in S. cerevisiae. The product of FOR1 has homology with a putative cell surface glycoprotein; FOR2 encodes glutamine:fructose-6-phosphate amidotransferase (GFAT), an enzyme that catalyzes the first step in the biosynthetic pathway leading to amino sugar-containing macromolecules, such as glycoproteins and chitin (64); and FOR3 has high homology with S. cerevisiae SSD1, a gene that controls cell wall composition and virulence (31, 78). FOR2 and FOR3 are the functional equivalents of the corresponding S. cerevisiae genes. Our parallel analysis using two model fungi verifies the notion that cell wall proteins play a critical role in determining the sensitivity/resistance of fungi to osmotin. In addition, these results implicate that the tobacco defense protein, osmotin, can serve as an effective/useful tool in identifying genes that control cell wall composition not only in a model fungus, such as S. cerevisiae, but also in phytopathogenic fungi, such as F. oxysporum.  相似文献   

4.
Tandem repeat (TR) regions are common in yeast adhesins, but their structures are unknown, and their activities are poorly understood. TR regions in Candida albicans Als proteins are conserved glycosylated 36-residue sequences with cell-cell aggregation activity (J. M. Rauceo, R. De Armond, H. Otoo, P. C. Kahn, S. A. Klotz, N. K. Gaur, and P. N. Lipke, Eukaryot. Cell 5:1664–1673, 2006). Ab initio modeling with either Rosetta or LINUS generated consistent structures of three-stranded antiparallel β-sheet domains, whereas randomly shuffled sequences with the same composition generated various structures with consistently higher energies. O- and N-glycosylation patterns showed that each TR domain had exposed hydrophobic surfaces surrounded by glycosylation sites. These structures are consistent with domain dimensions and stability measurements by atomic force microscopy (D. Alsteen, V. Dupres, S. A. Klotz, N. K. Gaur, P. N. Lipke, and Y. F. Dufrene, ACS Nano 3:1677–1682, 2009) and with circular dichroism determination of secondary structure and thermal stability. Functional assays showed that the hydrophobic surfaces of TR domains supported binding to polystyrene surfaces and other TR domains, leading to nonsaturable homophilic binding. The domain structures are like “classic” subunit interaction surfaces and can explain previously observed patterns of promiscuous interactions between TR domains in any Als proteins or between TR domains and surfaces of other proteins. Together, the modeling techniques and the supporting data lead to an approach that relates structure and function in many kinds of repeat domains in fungal adhesins.Yeast adhesins are a diverse set of cell adhesion proteins that mediate adhesion to host cells, environmental substrates, other fungi, and coinfecting bacteria (6, 8, 20, 21, 23, 29). The adhesins share common features, including compact N-terminal domains similar to Ig or lectin domains, Thr-rich midpieces, often in tandem repeats, and long highly glycosylated Ser/Thr-rich C-terminal regions that extend the functional domains out from the cell surface. No structures for the Thr-rich midpieces are known, but they can mediate aggregation of fungal cells (33, 35, 47). The prevalence and conservation of such repeats argue that they are functionally important, despite limited data on their structure and function.In Candida albicans, the Als adhesins are homologous proteins, products of 8 loci that encode numerous alleles of cell surface adhesins (16). In each mature Als protein, there are, from the N terminus, three tandem Ig-like domains, a β-sheet-rich conserved 127-residue amyloid-forming T region, a variable number of 36-residue tandem repeats (TRs), and a highly glycosylated stalk region that extends the N-terminal domains away from the cell surface (Fig. 1) (16, 33, 41). The C termini of these and other wall-associated adhesins are covalently cross-linked into the cell wall through transglycosylation of a modified glycosylphosphatidylinositol (GPI) anchor (18, 25). This modular design, including tandem repeats, is typical of fungal adhesins (8).Open in a separate windowFig. 1.Schematic diagram of the sequence of Als5p. The regions are named above, and the number of amino acid residues in each region is shown below. The modeled sequences are in the TR region.The Als protein Ig-like region, T region, and TR region all have protein-protein interaction activities (26, 33, 35). The Ig-like regions can interact with diverse mammalian proteins, presumably in a way analogous to antibody-antigen binding, as has been shown in the homologous protein α-agglutinin from Saccharomyces cerevisiae (8, 24, 26, 35). The T regions interact through formation of amyloid-like structures both in vivo and in vitro (33, 34a, 36). An insight into the function of the tandem repeats followed from observations that Als proteins initiate and maintain cell-to-cell aggregations, either spontaneously (“autoaggregation”) or following adhesion to a bead-bound defined ligand (10, 11, 36). Aggregation is more extensive for Als proteins with more tandem repeats (26, 35). This result suggested that the tandem repeats are uniquely structured to facilitate or mediate the aggregative function. Circular dichroism spectroscopy of the TR region of Als5p shows a β-sheet-rich structure in the soluble protein (35).In support of their direct involvement in aggregation, the repeat region of the C. albicans adhesin Als5p mediates cell-cell aggregation in the absence of the Ig-like and T domains (35). Moreover, the repeats can also potentiate binding of Als5p to fibronectin (35). Thus, the TR domains mediate cellular aggregation and increased binding to fibronectin. In addition, TR domains and their amino acid sequences are highly conserved across several Candida species (3). These properties need to be explained by their three-dimensional structure.Because there are no homologous structures known, we modeled by two independent ab initio methods. Rosetta assembles structures by combining short peptide structures extracted from the protein structural database PDB (38), then combines structures in a Monte Carlo approach, and assesses energetics of assembled structures. Rosetta has recently been shown to generate accurate models for protein-sized domains (40). We also predicted structures with LINUS, which generates randomized structures and rapidly estimates energetics to choose low-energy models (45). The models were supported by structural analyses with atomic force microscopy and circular dichroism spectroscopy. Functional assays showed that the TR domains can mediate binding activities predicted from the calculated structures.  相似文献   

5.
We are interested in asparagine-linked glycans (N-glycans) of Plasmodium falciparum and Toxoplasma gondii, because their N-glycan structures have been controversial and because we hypothesize that there might be selection against N-glycans in nucleus-encoded proteins that must pass through the endoplasmic reticulum (ER) prior to threading into the apicoplast. In support of our hypothesis, we observed the following. First, in protists with apicoplasts, there is extensive secondary loss of Alg enzymes that make lipid-linked precursors to N-glycans. Theileria makes no N-glycans, and Plasmodium makes a severely truncated N-glycan precursor composed of one or two GlcNAc residues. Second, secreted proteins of Toxoplasma, which uses its own 10-sugar precursor (Glc3Man5GlcNAc2) and the host 14-sugar precursor (Glc3Man9GlcNAc2) to make N-glycans, have very few sites for N glycosylation, and there is additional selection against N-glycan sites in its apicoplast-targeted proteins. Third, while the GlcNAc-binding Griffonia simplicifolia lectin II labels ER, rhoptries, and surface of plasmodia, there is no apicoplast labeling. Similarly, the antiretroviral lectin cyanovirin-N, which binds to N-glycans of Toxoplasma, labels ER and rhoptries, but there is no apicoplast labeling. We conclude that possible selection against N-glycans in protists with apicoplasts occurs by eliminating N-glycans (Theileria), reducing their length (Plasmodium), or reducing the number of N-glycan sites (Toxoplasma). In addition, occupation of N-glycan sites is markedly reduced in apicoplast proteins versus some secretory proteins in both Plasmodium and Toxoplasma.Animals, fungi, and plants synthesize Asn-linked glycans (N-glycans) by means of a lipid-linked precursor containing 14 sugars (dolichol-PP-Glc3Man9GlcNAc2) (26). Recently we used bioinformatics and experimental methods to show that numerous protists are missing sets of glycosyltransferases (Alg1 to Alg14) and so make truncated N-glycan precursors containing 0 to 11 sugars (46). For example, Entamoeba histolytica, which causes dysentery, makes N-glycan precursors that contain seven sugars (Man5GlcNAc2) (33). Giardia lamblia, a cause of diarrhea, makes N-glycan precursors that contain just GlcNAc2 (41). N-glycan precursors may be identified by metabolic labeling with radiolabeled mannose (Entamoeba) or glucosamine (Giardia) (46). Unprocessed N-glycans of each protist may be recognized by wheat germ agglutinin 1 (WGA-1) (GlcNAc2 of Giardia) or by the antiretroviral lectin cyanovirin-N (Man5GlcNAc2 of Entamoeba) (2, 33, 41).N-glycans are transferred from lipid-linked precursors to sequons (Asn-Xaa-Ser or Asn-Xaa-Thr, where Xaa cannot be Pro) on nascent peptides by an oligosaccharyltransferase (OST) (28). For the most part, transfer of N-glycans by the OST is during translocation, although there are human and Trypanosoma OSTs that transfer N-glycans after translocation (34, 45).N-glycan-dependent quality control (QC) systems for protein folding and endoplasmic reticulum (ER)-associated degradation (ERAD), which are present in most eukaryotes, are missing from Giardia and a few other protists that make truncated N-glycans (5, 26, 53). There is positive Darwinian selection for sequons (sites of N-glycans) that contain Thr in secreted and membrane proteins of organisms that have N-glycan-dependent QC (12). This selection occurs for the most part by an increased probability that Asn and Thr will be present in sequons rather than elsewhere in secreted and membrane proteins. In contrast, there is no selection on sequons that contain Ser, and there is no selection on sequons in the secreted proteins of organisms that lack N-glycan-dependent QC.For numerous reasons, we are interested in the N-glycans of Plasmodium falciparum and Toxoplasma gondii, which cause severe malaria and disseminated infections, respectively.(i) There has been controversy for a long time as to whether Plasmodium makes N-glycans. While some investigators identified a 14-sugar Plasmodium N-glycan resembling that of the human host (29), others identified no N-glycans (6, 22).(ii) There is also controversy concerning whether the N-glycans of Toxoplasma, after removal of Glc by glucosidases in the ER lumen, contain either 7 sugars (Man5GlcNAc2), like Entamoeba (32, 33), or 11 sugars (Man9GlcNAc2), like the human host (16, 19, 26). If it is Man5GlcNAc2, then Toxoplasma uses the dolichol-PP-linked glycan predicted by its set of Alg enzymes (32, 46). If it is Man9GlcNAc2, then Toxoplasma uses the dolichol-PP-linked glycan of the host cell (16, 19, 26).(iii) Both Plasmodium and Toxoplasma are missing proteins involved in N-glycan-dependent QC of protein folding (5).(iv) We hypothesize that there may be negative selection against N-glycans in Plasmodium and Toxoplasma, because the N-glycans added in the ER lumen during translocation will likely interfere with threading of nucleus-encoded apicoplast proteins into a nonphotosynthetic, chloroplast-derived organelle called the apicoplast (21, 35, 37, 48, 52, 54). Nucleus-encoded apicoplast proteins have a bipartite signal at the N terminus, which targets proteins first to the lumen of the ER and second to lumen of the apicoplast. This bipartite signal has been used in transformed plasmodia where green fluorescent protein (GFP) is targeted to the apicoplast with the bipartite signal of the acyl carrier protein (ACPleader-GFP), to the secretory system with the signal sequence only (ACPsignal-GFP), and to the cytosol with the organelle-targeting transit peptide only (ACPtransit-GFP) (55). Similar constructs have been used to characterize signals that target nucleus-encoded proteins of Toxoplasma to the apicoplast (11, 25).Here we use a combination of bioinformatic, biochemical, and morphological methods to characterize the N-glycans of Plasmodium and Toxoplasma and to test our hypothesis that there is negative selection against N-glycans in protists with apicoplasts.  相似文献   

6.
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   

7.
Two methods were used to compare the biodegradation of six polychlorinated biphenyl (PCB) congeners by 12 white rot fungi. Four fungi were found to be more active than Phanerochaete chrysosporium ATCC 24725. Biodegradation of the following congeners was monitored by gas chromatography: 2,3-dichlorobiphenyl, 4,4′-dichlorobiphenyl, 2,4′,5-trichlorobiphenyl (2,4′,5-TCB), 2,2′,4,4′-tetrachlorobiphenyl, 2,2′,5,5′-tetrachlorobiphenyl, and 2,2′,4,4′,5,5′-hexachlorobiphenyl. The congener tested for mineralization was 2,4′,5-[U-14C]TCB. Culture supernatants were also assayed for lignin peroxidase and manganese peroxidase activities. Of the fungi tested, two strains of Bjerkandera adusta (UAMH 8258 and UAMH 7308), one strain of Pleurotus ostreatus (UAMH 7964), and Trametes versicolor UAMH 8272 gave the highest biodegradation and mineralization. P. chrysosporium ATCC 24725, a strain frequently used in studies of PCB degradation, gave the lowest mineralization and biodegradation activities of the 12 fungi reported here. Low but detectable levels of lignin peroxidase and manganese peroxidase activity were present in culture supernatants, but no correlation was observed among any combination of PCB congener biodegradation, mineralization, and lignin peroxidase or manganese peroxidase activity. With the exception of P. chrysosporium, congener loss ranged from 40 to 96%; however, these values varied due to nonspecific congener binding to fungal biomass and glassware. Mineralization was much lower, ≤11%, because it measures a complete oxidation of at least part of the congener molecule but the results were more consistent and therefore more reliable in assessment of PCB biodegradation.

Polychlorinated biphenyls (PCBs) are produced by chlorination of biphenyl, resulting in up to 209 different congeners. Commercial mixtures range from light oily fluids to waxes, and their physical properties make them useful as heat transfer fluids, hydraulic fluids, solvent extenders, plasticizers, flame retardants, organic diluents, and dielectric fluids (1, 21). Approximately 24 million lb are in the North American environment (19). The stability and hydrophobic nature of these compounds make them a persistent environmental hazard.To date, bacterial transformations have been the main focus of PCB degradation research. Aerobic bacteria use a biphenyl-induced dioxygenase enzyme system to attack less-chlorinated congeners (mono- to hexachlorobiphenyls) (1, 5, 7, 8, 22). Although more-chlorinated congeners are recalcitrant to aerobic bacterial degradation, microorganisms in anaerobic river sediments reductively dechlorinate these compounds, mainly removing the meta and para chlorines (1, 6, 10, 33, 34).The degradation of PCBs by white rot fungi has been known since 1985 (11, 18). Many fungi have been tested for their ability to degrade PCBs, including the white rot fungi Coriolus versicolor (18), Coriolopsis polysona (41), Funalia gallica (18), Hirneola nigricans (35), Lentinus edodes (35), Phanerochaete chrysosporium (3, 11, 14, 17, 18, 35, 39, 4143), Phlebia brevispora (18), Pleurotus ostreatus (35, 43), Poria cinerescens (18), Px strain (possibly Lentinus tigrinus) (35), and Trametes versicolor (41, 43). There have also been studies of PCB metabolism by ectomycorrhizal fungi (17) and other fungi such as Aspergillus flavus (32), Aspergillus niger (15), Aureobasidium pullulans (18), Candida boidinii (35), Candida lipolytica (35), Cunninghamella elegans (16), and Saccharomyces cerevisiae (18, 38). The mechanism of PCB biodegradation has not been definitively determined for any fungi. White rot fungi produce several nonspecific extracellular enzymes which have been the subject of extensive research. These nonspecific peroxidases are normally involved in lignin degradation but can oxidize a wide range of aromatic compounds including polycyclic aromatic hydrocarbons (37). Two peroxidases, lignin peroxidase (LiP) and Mn peroxidase (MnP), are secreted into the environment of the fungus under conditions of nitrogen limitation in P. chrysosporium (23, 25, 27, 29) but are not stress related in fungi such as Bjerkandera adusta or T. versicolor (12, 30).Two approaches have been used to determine the biodegradability of PCBs by fungi: (i) loss of the parent congener analyzed by gas chromatography (GC) (17, 32, 35, 42, 43) and (ii) mineralization experiments in which the 14C of the universally labeled 14C parent congener is recovered as 14CO2 (11, 14, 18, 39, 41). In the first method, the loss of a peak on a chromatogram makes it difficult to decide whether the PCB is being partly degraded, mineralized, adsorbed to the fungal biomass, or bound to glassware, soil particles, or wood chips. Even when experiments with killed-cell and abiotic controls are performed, the extraction efficiency and standard error can make data difficult to interpret. For example, recoveries can range anywhere from 40 to 100% depending on the congener used and the fungus being investigated (17). On the other hand, recovery of significant amounts of 14CO2 from the cultures incubated with a 14C substrate provides definitive proof of fungal metabolism. There appears to be only one report relating data from these two techniques (18), and in that study, [U-14C]Aroclor 1254, rather than an individual congener, was used.In this study, we examined the ability of 12 white rot fungal strains to metabolize selected PCB congeners to determine which strains were the most active degraders. Included in this group was P. chrysosporium ATCC 24725, a strain used extensively in PCB studies (3, 14, 18, 35, 39, 4143). Six PCB congeners were selected to give a range of chlorine substitutions and therefore a range of potential biodegradability which was monitored by GC. One of the chosen congeners was 14C labeled and used in studies to compare the results from a mineralization method with those from the GC method.  相似文献   

8.
9.
The opportunistic human pathogen Acinetobacter baumannii is a concern to health care systems worldwide because of its persistence in clinical settings and the growing frequency of multiple drug resistant infections. To combat this threat, it is necessary to understand factors associated with disease and environmental persistence of A. baumannii. Recently, it was shown that a single biosynthetic pathway was responsible for the generation of capsule polysaccharide and O-linked protein glycosylation. Because of the requirement of these carbohydrates for virulence and the non-template driven nature of glycan biogenesis we investigated the composition, diversity, and properties of the Acinetobacter glycoproteome. Utilizing global and targeted mass spectrometry methods, we examined 15 strains and found extensive glycan diversity in the O-linked glycoproteome of Acinetobacter. Comparison of the 26 glycoproteins identified revealed that different A. baumannii strains target similar protein substrates, both in characteristics of the sites of O-glycosylation and protein identity. Surprisingly, glycan micro-heterogeneity was also observed within nearly all isolates examined demonstrating glycan heterogeneity is a widespread phenomena in Acinetobacter O-linked glycosylation. By comparing the 11 main glycoforms and over 20 alternative glycoforms characterized within the 15 strains, trends within the glycan utilized for O-linked glycosylation could be observed. These trends reveal Acinetobacter O-linked glycosylation favors short (three to five residue) glycans with limited branching containing negatively charged sugars such as GlcNAc3NAcA4OAc or legionaminic/pseudaminic acid derivatives. These observations suggest that although highly diverse, the capsule/O-linked glycan biosynthetic pathways generate glycans with similar characteristics across all A. baumannii.Acinetobacter baumannii is an emerging opportunistic pathogen of increasing significance to health care institutions worldwide (13). The growing number of identified multiple drug resistant (MDR)1 strains (24), the ability of isolates to rapidly acquire resistance (3, 4), and the propensity of this agent to survive harsh environmental conditions (5) account for the increasing number of outbreaks in intensive care, burn, or high dependence health care units since the 1970s (25). The burden on the global health care system of MDR A. baumannii is further exacerbated by standard infection control measures often being insufficient to quell the spread of A. baumannii to high risk individuals and generally failing to remove A. baumannii from health care institutions (5). Because of these concerns, there is an urgent need to identify strategies to control A. baumannii as well as understand the mechanisms that enable its persistence in health care environments.Surface glycans have been identified as key virulence factors related to persistence and virulence within the clinical setting (68). Acinetobacter surface carbohydrates were first identified and studied in A. venetianus strain RAG-1, leading to the identification of a gene locus required for synthesis and export of the surface carbohydrates (9, 10). These carbohydrate synthesis loci are variable yet ubiquitous in A. baumannii (11, 12). Comparison of 12 known capsule structures from A. baumannii with the sequences of their carbohydrate synthesis loci has provided strong evidence that these loci are responsible for capsule synthesis with as many as 77 distinct serotypes identified by molecular serotyping (11). Because of the non-template driven nature of glycan synthesis, the identification and characterization of the glycans themselves are required to confirm the true diversity. This diversity has widespread implications for Acinetobacter biology as the resulting carbohydrate structures are not solely used for capsule biosynthesis but can be incorporated and utilized by other ubiquitous systems, such as O-linked protein glycosylation (13, 14).Although originally thought to be restricted to species such as Campylobacter jejuni (15, 16) and Neisseria meningitidis (17), bacterial protein glycosylation is now recognized as a common phenomenon within numerous pathogens and commensal bacteria (18, 19). Unlike eukaryotic glycosylation where robust and high-throughput technologies now exist to enrich (2022) and characterize both the glycan and peptide component of glycopeptides (2325), the diversity (glycan composition and linkage) within bacterial glycosylation systems makes few technologies broadly applicable to all bacterial glycoproteins. Because of this challenge a deeper understanding of the glycan diversity and substrates of glycosylation has been largely unachievable for the majority of known bacterial glycosylation systems. The recent implementation of selective glycopeptide enrichment methods (26, 27) and the use of multiple fragmentation approaches (28, 29) has facilitated identification of an increasing number of glycosylation substrates independent of prior knowledge of the glycan structure (3033). These developments have facilitated the undertaking of comparative glycosylation studies, revealing glycosylation is widespread in diverse genera and far more diverse then initially thought. For example, Nothaft et al. were able to show N-linked glycosylation was widespread in the Campylobacter genus and that two broad groupings of the N-glycans existed (34).During the initial characterization of A. baumannii O-linked glycosylation the use of selective enrichment of glycopeptides followed by mass spectrometry analysis with multiple fragmentation technologies was found to be an effective means to identify multiple glycosylated substrates in the strain ATCC 17978 (14). Interestingly in this strain, the glycan utilized for protein modification was identical to a single subunit of the capsule (13) and the loss of either protein glycosylation or glycan synthesis lead to decreases in biofilm formation and virulence (13, 14). Because of the diversity in the capsule carbohydrate synthesis loci and the ubiquitous distribution of the PglL O-oligosaccharyltransferase required for protein glycosylation, we hypothesized that the glycan variability might be also extended to O-linked glycosylation. This diversity, although common in surface carbohydrates such as the lipopolysaccharide of numerous Gram-negative pathogens (35), has only recently been observed within bacterial proteins glycosylation system that are typically conserved within species (36) and loosely across genus (34, 37).In this study, we explored the diversity within the O-linked protein glycosylation systems of Acinetobacter species. Our analysis complements the recent in silico studies of A. baumannii showing extensive glycan diversity exists in the carbohydrate synthesis loci (11, 12). Employing global strategies for the analysis of glycosylation, we experimentally demonstrate that the variation in O-glycan structure extends beyond the genetic diversity predicted by the carbohydrate loci alone and targets proteins of similar properties and identity. Using this knowledge, we developed a targeted approach for the detection of protein glycosylation, enabling streamlined analysis of glycosylation within a range of genetic backgrounds. We determined that; O-linked glycosylation is widespread in clinically relevant Acinetobacter species; inter- and intra-strain heterogeneity exist within glycan structures; glycan diversity, although extensive results in the generation of glycans with similar properties and that the utilization of a single glycan for capsule and O-linked glycosylation is a general feature of A. baumannii but may not be a general characteristic of all Acinetobacter species such as A. baylyi.  相似文献   

10.
11.
12.
13.
14.
15.
The intermembrane space (IMS) represents the smallest subcompartment of mitochondria. Nevertheless, it plays important roles in the transport and modification of proteins, lipids, and metal ions and in the regulation and assembly of the respiratory chain complexes. Moreover, it is involved in many redox processes and coordinates key steps in programmed cell death. A comprehensive profiling of IMS proteins has not been performed so far. We have established a method that uses the proapoptotic protein Bax to release IMS proteins from isolated mitochondria, and we profiled the protein composition of this compartment. Using stable isotope-labeled mitochondria from Saccharomyces cerevisiae, we were able to measure specific Bax-dependent protein release and distinguish between quantitatively released IMS proteins and the background efflux of matrix proteins. From the known 31 soluble IMS proteins, 29 proteins were reproducibly identified, corresponding to a coverage of >90%. In addition, we found 20 novel intermembrane space proteins, out of which 10 had not been localized to mitochondria before. Many of these novel IMS proteins have unknown functions or have been reported to play a role in redox regulation. We confirmed IMS localization for 15 proteins using in organello import, protease accessibility upon osmotic swelling, and Bax-release assays. Moreover, we identified two novel mitochondrial proteins, Ymr244c-a (Coa6) and Ybl107c (Mic23), as substrates of the MIA import pathway that have unusual cysteine motifs and found the protein phosphatase Ptc5 to be a novel substrate of the inner membrane protease (IMP). For Coa6 we discovered a role as a novel assembly factor of the cytochrome c oxidase complex. We present here the first and comprehensive proteome of IMS proteins of yeast mitochondria with 51 proteins in total. The IMS proteome will serve as a valuable source for further studies on the role of the IMS in cell life and death.Mitochondria are double-membrane-bound organelles that fulfill a multitude of important cellular functions. Proteomic analysis of purified mitochondria revealed that they contain approximately 1000 (yeast) to 1500 (human) different proteins (13). However, the distribution of these proteins among the four mitochondrial subcompartments (outer membrane, inner membrane, matrix, and intermembrane space) has been only marginally studied through global approaches. This is attributed to the high complexity of purifying submitochondrial fractions to a grade suitable for proteomic analysis. The best-studied submitochondrial proteomes comprise the outer membranes of S. cerevisae, N. crassa, and A. thaliana (46). The mitochondrial intermembrane space (IMS)1 represents a highly interesting compartment for several reasons: it provides a redox active space that promotes oxidation of cysteine residues similar to the endoplasmic reticulum and the bacterial periplasm, but unlike cytosol, nucleus, or the mitochondrial matrix where the presence of thioredoxins or glutaredoxins prevents the risk of unwanted cysteine oxidation (7, 8). Furthermore in higher eukaryotes IMS proteins are released into the cytosol upon apoptotic induction, which triggers the activation of a cell-killing protease activation cascade (9, 10). The IMS can also exchange proteins, lipids, metal ions, and various metabolites with other cellular compartments, allowing mitochondrial metabolism to adapt to cellular homeostasis. In particular, the biogenesis and activity of the respiratory chain were shown to be controlled by various proteins of the IMS (1113). Most of the currently known IMS proteins are soluble proteins; however, some inner membrane proteins have been annotated as IMS proteins as well, such as proteins that are peripherally attached to the inner membrane or membrane proteins that expose enzyme activity toward the IMS (8).All IMS proteins are encoded in the nuclear DNA and have to be imported after translation in the cytosol (1419). Two main pathways are known to mediate the import and sorting of proteins into the IMS. One class of proteins contains bipartite presequences that consist of a matrix targeting signal and a hydrophobic sorting signal. These signals arrest the incoming preprotein at the inner membrane translocase TIM23. After insertion into the inner membrane, the soluble, mature protein can be released into the IMS by the inner membrane protease (IMP) (2022). The second class of IMS proteins possesses characteristic cysteine motifs that typically are either twin CX9C or twin CX3C motifs (23, 24). Upon translocation across the outer membrane via the TOM complex, disulfide bonds are formed within the preproteins, which traps them in the IMS. Disulfide bond formation is mediated by the MIA machinery, which consists of the inner-membrane-anchored Mia40 and the soluble IMS protein Erv1 (2528).The release of cytochrome c from the IMS upon binding and insertion of Bax at the outer membrane is a hallmark of programmed cell death. Although Bax is found only in higher eukaryotes, it was shown that recombinant mammalian Bax induces the release of cytochrome c upon incubation with isolated yeast mitochondria (29, 30). Furthermore, we found that not only cytochrome c but also other soluble IMS proteins are released from Bax-treated yeast mitochondria, whereas soluble matrix proteins largely remain within the organelle (30).We used this apparently conserved mechanism to systematically profile the protein composition of the yeast mitochondrial IMS by employing an experimental approach based on stable isotope labeling, which allowed for the specific identification of Bax-dependent protein release. Almost the entire set of known soluble IMS proteins was identified, and 20 additional, novel soluble IMS proteins were found. We confirmed IMS localization for 15 proteins through biochemical assays. Among these proteins, we identified novel proteins that fall into several classes: (i) those that are involved in maintaining protein redox homeostasis (thioredoxins, thioredoxin reductases, or thiol peroxidases), (ii) those that undergo proteolytic processing by IMP (Ptc5), (iii) those that utilize the MIA pathway for their import (Mic23 and Coa6), and (iv) those that play a role in the assembly of cytochrome c oxidase (Coa6).  相似文献   

16.
The accumulation of bioenergy carriers was assessed in two starchless mutants of Chlamydomonas reinhardtii (the sta6 [ADP-glucose pyrophosphorylase] and sta7-10 [isoamylase] mutants), a control strain (CC124), and two complemented strains of the sta7-10 mutant. The results indicate that the genetic blockage of starch synthesis in the sta6 and sta7-10 mutants increases the accumulation of lipids on a cellular basis during nitrogen deprivation relative to that in the CC124 control as determined by conversion to fatty acid methyl esters. However, this increased level of lipid accumulation is energetically insufficient to completely offset the loss of cellular starch that is synthesized by CC124 during nitrogen deprivation. We therefore investigated acetate utilization and O2 evolution to obtain further insights into the physiological adjustments utilized by the two starchless mutants in the absence of starch synthesis. The results demonstrate that both starchless mutants metabolize less acetate and have more severely attenuated levels of photosynthetic O2 evolution than CC124, indicating that a decrease in overall anabolic processes is a significant physiological response in the starchless mutants during nitrogen deprivation. Interestingly, two independent sta7-10:STA7 complemented strains exhibited significantly greater quantities of cellular starch and lipid than CC124 during acclimation to nitrogen deprivation. Moreover, the complemented strains synthesized significant quantities of starch even when cultured in nutrient-replete medium.Microalgae are able to efficiently convert sunlight, water, and CO2 into a variety of products suitable for renewable energy applications, including H2, carbohydrates, and lipids (11, 12, 16, 21, 38, 41, 44). The unicellular green alga Chlamydomonas reinhardtii has emerged as a model organism for studying algal physiology, photosynthesis, metabolism, nutrient stress, and the synthesis of bioenergy carriers (12, 15, 19, 24, 32). During acclimation to nitrogen deprivation, C. reinhardtii cells accumulate significant quantities of starch and form lipid bodies (4, 5, 8, 26, 28, 30, 34, 43, 46, 48). Despite the significance of these products in algal physiology and in biofuels applications, the metabolic, enzymatic, and regulatory mechanisms controlling the partitioning of metabolites into these distinct carbon stores in algae are poorly understood. Several C. reinhardtii starch mutants with various phenotypic changes in starch content and structure have been isolated (2,4). Two of these, the sta6 and sta7 mutants, contain single-gene disruptions that result in “starchless” phenotypes with severely attenuated levels of starch granule accumulation (2, 4, 34, 39, 40, 48).The disrupted loci in the two isolated starchless mutants are distinct and each mutant has a unique phenotype (7, 40). In the sta6 mutant, the small, catalytic subunit of ADP-glucose pyrophosphorylase (AGPase-SS) is disrupted (2, 4, 48), and this mutant accumulates less than 1% of the starch observed in wild-type (WT) cells under conditions of nitrogen deprivation. The sta7 mutant contains a disrupted isoamylase gene (7, 8, 10, 39, 40) and also has severely attenuated levels of starch, but it accumulates a soluble glycogen-like product (4, 9). In this study, we conducted an examination of the unique physiological acclimations that are utilized by these mutants to adapt to the loss of starch synthesis. As the genetic lesions in these two mutants are distinct and block starch synthesis via two very different mechanisms, we investigated the physiological consequences of starch inhibition in both of these mutants from a holistic bioenergy perspective, which included photosynthetic parameters and the overall yields of lipids and carbohydrates, the two primary bioenergy carriers in C. reinhardtii. Specifically, we examined whether the inability to synthesize starch would result in the accumulation of additional lipid, alter cellular growth or cell size, affect acetate utilization, and/or influence photosynthetic O2 evolution. Our data indicate that both the sta6 (BAFJ5) and sta7 (sta7-10) mutants accumulate more lipid than the CC124 control during nitrogen deprivation. However, the additional lipid does not completely offset the loss of starch synthesis from a complete energetic perspective. Increased lipid accumulation during nitrogen stress has also been reported for a variety of starch mutants in recent papers (26, 27, 46). A significant feature in both of the starchless mutants studied here is that O2 evolution and acetate utilization are diminished during nitrogen stress, which is undesirable from an overall bioenergy perspective. Remarkably, complementation of sta7-10 with genomic DNA encoding the wild-type isoamylase gene resulted in cells that were larger than those of the sta6, sta7-10, and CC124 strains, exhibited the highest total lipid levels during nitrogen deprivation, and overaccumulated starch even in nutrient-replete medium.  相似文献   

17.
The occurrence of highly conserved amyloid-forming sequences in Candida albicans Als proteins (H. N. Otoo et al., Eukaryot. Cell 7:776–782, 2008) led us to search for similar sequences in other adhesins from C. albicans and Saccharomyces cerevisiae. The β-aggregation predictor TANGO found highly β-aggregation-prone sequences in almost all yeast adhesins. These sequences had an unusual amino acid composition: 77% of their residues were β-branched aliphatic amino acids Ile, Thr, and Val, which is more than 4-fold greater than their prevalence in the S. cerevisiae proteome. High β-aggregation potential peptides from S. cerevisiae Flo1p and C. albicans Eap1p rapidly formed insoluble amyloids, as determined by Congo red absorbance, thioflavin T fluorescence, and fiber morphology. As examples of the amyloid-forming ability of the native proteins, soluble glycosylphosphatidylinositol (GPI)-less fragments of C. albicans Als5p and S. cerevisiae Muc1p also formed amyloids within a few days under native conditions at nM concentrations. There was also evidence of amyloid formation in vivo: the surfaces of cells expressing wall-bound Als1p, Als5p, Muc1p, or Flo1p were birefringent and bound the fluorescent amyloid-reporting dye thioflavin T. Both of these properties increased upon aggregation of the cells. In addition, amyloid binding dyes strongly inhibited aggregation and flocculation. The results imply that amyloid formation is an intrinsic property of yeast cell adhesion proteins from many gene families and that amyloid formation is an important component of cellular aggregation mediated by these proteins.Protein amyloids are characteristic of pathological conditions, including neurodegenerative diseases (4, 11, 17, 38). These protein aggregates can also occur naturally in adhesive bacterial curli (3), melanosomes (14), condensed peptide hormone arrays (24), as regulatory prions in yeast (2, 5), and fungal hydrophobins, which are nonantigenic coats to some fungi (1, 33, 39). Nevertheless, such natural occurrences are relatively few, considering the negative free energy for amyloid formation (28).We have recently discovered that there are amyloid-forming sequences in the cell surface Als adhesins of Candida albicans. Cells that express these adhesins aggregate readily, and the aggregation has amyloid-like properties, including protein conformational shifting, surface birefringence, and ability to bind the amyloid-active dyes Congo red and amino-naphthalene sulfonic acid (ANS) (29). A five- to seven-residue sequence in Als1p, Als3p, and Als5p has extremely high potential for formation of β-aggregates, according to the protein state prediction program TANGO (13, 27, 31). Such β-aggregates include amyloids, which are ordered structures with paracrystalline regions of stacked parallel β-strands that are perpendicular to the long axis of micrometer-long fibrils. The strands are stabilized by interaction of identical sequences from many protein molecules (31, 32). Where TANGO analyses have shown that specific sequences have β-aggregate potentials greater than 20%, an insoluble β-aggregate state is likely to form. These β-aggregates nucleate formation of amyloids if the proteins can associate to form fibers (13, 27, 31). Sequences in the conserved 127-residue T region of Als1p, Als3p, and Als5p have β-aggregation potentials of >90% (27). An oligopeptide with this sequence, as well as 412- and 645-residue fragments of Als5p formed authentic amyloids, as determined by characteristic dye binding and fiber morphology. The amyloid-forming sequences were rich in the β-branched amino acids Thr, Val, and Ile. This amino acid composition is unusual among proteins in general, but is common in the Thr-rich mid-piece domains of yeast adhesins.Yeasts display many cell-wall-bound adhesins that mediate colonial and biofilm interactions as well as host-pathogen binding (9, 21, 41). Such adhesins have a common mosaic structure. In general, the adhesins have N-terminal globular binding domains (often immunoglobulin-like or lectin-like), Thr-rich mid-piece sequences including tandem repeats, and 300- to 800-residue heavily glycosylated Ser and Thr-rich “stalk” domains near the C-terminal domain that extend the active regions from the surface of the wall. The adhesins are covalently cross-linked to wall polysaccharides through modified glycosylphosphatidylinositol (GPI) anchors and/or glycosyl esters of glutamic acid (9, 18).Because the yeast adhesins share this common modular domain structure, we searched among known and putative yeast adhesins for sequences with high β-aggregation potential. We have found that many of these proteins share amyloid-forming sequences and amyloid-like behavior on activation.  相似文献   

18.
19.
In many fungal pathogens, infection is initiated by conidial germination. Subsequent stages involve germ tube elongation, conidiation, and vegetative hyphal fusion (anastomosis). Here, we used live-cell fluorescence to study the dynamics of green fluorescent protein (GFP)- and cherry fluorescent protein (ChFP)-labeled nuclei in the plant pathogen Fusarium oxysporum. Hyphae of F. oxysporum have uninucleated cells and exhibit an acropetal nuclear pedigree, where only the nucleus in the apical compartment is mitotically active. In contrast, conidiation follows a basopetal pattern, whereby mononucleated microconidia are generated by repeated mitotic cycles of the subapical nucleus in the phialide, followed by septation and cell abscission. Vegetative hyphal fusion is preceded by directed growth of the fusion hypha toward the receptor hypha and followed by a series of postfusion nuclear events, including mitosis of the apical nucleus of the fusion hypha, migration of a daughter nucleus into the receptor hypha, and degradation of the resident nucleus. These previously unreported patterns of nuclear dynamics in F. oxysporum could be intimately related to its pathogenic lifestyle.Fusarium oxysporum is a soilborne pathogen that causes substantial losses in a wide variety of crops (12) and has been reported as an emerging human pathogen (36, 38). Similar to other fungal pathogens (18), the early stages of interaction between F. oxysporum and the host are crucial for the outcome of infection (11). Key processes occurring during these initial stages include spore germination, adhesion to the host surface, establishment of hyphal networks through vegetative hyphal fusion, differentiation of infection hyphae, and penetration of the host (53). Surprisingly, very little is known about the cytology of basic processes, such as spore germination and hyphal development, which play key roles during infection by F. oxysporum.F. oxysporum produces three types of asexual spores: microconidia, macroconidia, and chlamydospores (9, 26). Germination usually represents the first step in the colonization of a new environment, including the host. Once dormancy is broken, spores undergo a defined set of morphogenetic changes that lead to the establishment of a polarized growth axis and the emergence of one or multiple germ tubes (reviewed by d''Enfert and Hardham [10, 19]). In certain fungi, such as Aspergillus nidulans, germ tube emergence and septum formation are subject to precise spatial controls and are tightly coordinated with nuclear division (20, 22, 34, 42, 54). In contrast, in spores from other filamentous fungi, such as macroconidia of Fusarium graminearum, nuclear division is not required for the emergence of germ tubes (21, 48). During hyphal growth, multinucleate fungi display distinct mitotic patterns, such as asynchronous nuclear division in Neurospora crassa and Ashbya gossypii (15, 16, 29, 30, 33, 49), parasynchronous in A. nidulans (7, 15, 23, 46), and synchronous in Ceratocystis fagacearum (1, 15).Vegetative hyphal fusion, or anastomosis, is a common developmental process during the life cycle of filamentous fungi that is thought to serve important functions in intrahyphal communication, nutrient transport, and colony homeostasis (41). F. oxysporum undergoes anastomosis (8, 25, 32, 40), and although this process is not strictly required for plant infection, it appears to contribute to efficient colonization of the root surface (39).The aim of this study was to explore nuclear dynamics during different developmental stages of F. oxysporum that are of key relevance during the establishment of infection. They include germination of microconidia, vegetative hyphal development, and conidiation, as well as vegetative hyphal fusion during colony establishment. Fusion PCR-mediated gene targeting (55) was used to C-terminally label histone H1 in F. oxysporum (FoH1) with either green fluorescent protein (GFP) or the cherry variant (ChFP), allowing us to perform, for the first time, live-cell analysis of nuclear dynamics in this species. Our study revealed distinct patterns of nuclear divisions in F. oxysporum. Moreover, we report, for the first time in an ascomycete, that hyphal fusion initiates a series of nuclear events, including mitosis in the fusing hypha and nuclear migration into the receptor hypha, followed by degradation of the resident nucleus.  相似文献   

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