首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
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   

3.
4.
5.
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   

6.
7.
Filamentous fungi synthesize bioactive secondary metabolites with major human health and economic impacts. Little is known about the mechanisms that mediate the export of these metabolites to the cell exterior. Aspergillus parasiticus synthesizes aflatoxin, a secondary metabolite that is one of the most potent naturally occurring carcinogens known. We previously demonstrated that aflatoxin is synthesized and compartmentalized in specialized vesicles called aflatoxisomes and that these subcellular organelles also play a role in the export process. In the current study, we tested the hypothesis that aflatoxisomes fuse with the cytoplasmic membrane to facilitate the release of aflatoxin into the growth environment. Microscopic analysis of A. parasiticus grown under aflatoxin-inducing and non-aflatoxin-inducing conditions generated several lines of experimental evidence that supported the hypothesis. On the basis of the evidence, we propose that export of the mycotoxin aflatoxin in Aspergillus parasiticus occurs by exocytosis, and we present a model to illustrate this export mechanism.Secondary metabolites are chemically diverse natural products synthesized by plants, fungi, bacteria, algae, and animals. Secondary metabolites have an enormous impact on humans. Antibiotics, for example, are essential elements of the multibillion-dollar pharmaceutical industry, whereas mycotoxins cause hundreds of millions of dollars in damage to agriculture annually (11, 15). These chemicals help the producing organism to survive nutrient limitation (16). They also contribute to cellular defense mechanisms and development (11, 12), reduce cellular oxidative stress (10), and help maintain cellular homeostasis by regulating carbon flow in the cell (17).Many fungal secondary metabolites are exported outside the cell; examples include antibiotics and mycotoxins (3, 14). We and others conducted extensive studies on the regulation of fungal secondary metabolism at the molecular (11, 15) and cellular (3, 7) levels. However, little is known about the mechanisms that mediate secondary metabolite export or why export occurs.The filamentous fungus Aspergillus parasiticus produces aflatoxin, a secondary metabolite and the most potent naturally occurring carcinogen known. More than 90% of aflatoxin is exported to the cell exterior (3), making A. parasiticus an excellent model for studying secondary metabolite export. We recently demonstrated that specialized trafficking vesicles called aflatoxisomes play a key role in aflatoxin synthesis and export (3). As synthesis initiates, vesicle-vacuole fusion is downregulated by the global regulator Velvet, resulting in the accumulation of aflatoxisomes which contain at least the last two functional enzymes in the aflatoxin pathway and sequester aflatoxin (3). Treatments that block vesicle-vacuole fusion increase the number of aflatoxisomes, increase the quantity of aflatoxin accumulated in aflatoxisomes, and increase aflatoxin export to the cell exterior (3). On the basis of these previous observations, we hypothesized that aflatoxisomes play a direct role in aflatoxin export.Vesicle-mediated export could theoretically occur by one (or more) of at least three mechanisms (Fig. 1). (i) Vesicles pass across the cytoplasmic membrane intact and “shuttle” their contents into the external environment. This proposed mechanism mediates virulence factor release in Cryptococcus neoformans and Histoplasma capsulatum (1) during pathogenesis. (ii) Vesicles fuse to the cytoplasmic membrane and “pump” vesicle contents to the exterior using transporter proteins similar to those that mediate resistance to antifungal agents (4, 5). (iii) Vesicles fuse with the cytoplasmic membrane, which evaginates, bursts, and “blasts” vesicle contents to the exterior. This process is similar to exocytosis, a proposed secretory mechanism for specific proteins in filamentous fungi (18). We conducted the current study to determine which, if any, of these possible mechanisms most accurately reflects the process of aflatoxin export in A. parasiticus.Open in a separate windowFig. 1.Theoretical models for vesicle-mediated export. Aflatoxigenic vesicles (aflatoxisomes) arise due to downregulation of tethering complex (Tc) activity mediated by VeA (1). Aflatoxin synthesized in aflatoxisomes could theoretically be released to the cell exterior by one or more of three mechanisms: the shuttle (in which aflatoxisomes shuttle cargo across cytoplasmic membrane), pump (in which transmembrane transporter [Tp] proteins mediate the release of secondary metabolites as vesicles adhere to the inner surface of the cytoplasmic membrane), and burst-and-blast (in which vesicles protrude from the cell surface and blast their cargo into the medium) mechanisms. PM, plasma membrane.  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.
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   

14.
15.
16.
Identifying the sulfenylation state of stressed cells is emerging as a strategic approach for the detection of key reactive oxygen species signaling proteins. Here, we optimized an in vivo trapping method for cysteine sulfenic acids in hydrogen peroxide (H2O2) stressed plant cells using a dimedone based DYn-2 probe. We demonstrated that DYn-2 specifically detects sulfenylation events in an H2O2 dose- and time-dependent way. With mass spectrometry, we identified 226 sulfenylated proteins after H2O2 treatment of Arabidopsis cells, residing in the cytoplasm (123); plastid (68); mitochondria (14); nucleus (10); endoplasmic reticulum, Golgi and plasma membrane (7) and peroxisomes (4). Of these, 123 sulfenylated proteins have never been reported before to undergo cysteine oxidative post-translational modifications in plants. All in all, with this DYn-2 approach, we have identified new sulfenylated proteins, and gave a first glance on the locations of the sulfenomes of Arabidopsis thaliana.Among the different amino acids, the sulfur containing amino acids like cysteine are particularly susceptible to oxidation by reactive oxygen species (ROS)1 (1, 2). Recent studies suggest that the sulfenome, the initial oxidation products of cysteine residues, functions as an intermediate state of redox signaling (3 5). Thus, identifying the sulfenome under oxidative stress is a way to detect potential redox sensors (6, 7).This central role of the sulfenome in redox signaling provoked chemical biologists to develop strategies for sensitive detection and identification of sulfenylated proteins. The in situ trapping of the sulfenome is challenging because of two major factors: (1) the highly reactive, transient nature of sulfenic acids, which might be over-oxidized in excess of ROS, unless immediately protected by disulfide formation (7); (2) the intracellular compartmentalization of the redox state that might be disrupted during extraction procedures, resulting in artificial non-native protein oxidations (8, 9). Having a sulfur oxidation state of zero, sulfenic acids can react as both electrophile and nucleophile, however, direct detection methods are based on the electrophilic character of sulfenic acid (10). In 1974, Allison and coworkers reported a condensation reaction between the electrophilic sulfenic acid and the nucleophile dimedone (5,5-dimethyl-1,3-cyclohexanedione), producing a corresponding thioether derivative (11). This chemistry is highly selective and, since then, has been exploited to detect dimedone modified sulfenic acids using mass spectrometry (12). However, dimedone has limited applications for cellular sulfenome identification because of the lack of a functional group to enrich the dimedone tagged sulfenic acids. Later, dimedone-biotin/fluorophores conjugates have been developed, which allowed sensitive detection and enrichment of sulfenic acid modified proteins (13 15). This approach, however, was not always compatible with in vivo cellular sulfenome analysis, because the biotin/fluorophores-conjugated dimedone is membrane impermeable (9) and endogenous biotinylated proteins might appear as false positives.More recently, the Carroll lab has developed DYn-2, a sulfenic acid specific chemical probe. This chemical probe consists of two functional units: a dimedone scaffold for sulfenic acid recognition and an alkyne chemical handle for enrichment of labeled proteins (9). Once the sulfenic acids are tagged with the DYn-2 probe, they can be biotinylated through click chemistry (16). The click reaction used here is a copper (I)-catalyzed azide-alkyne cycloaddition reaction (17), also known as azide-alkyne Huisgen cycloaddition (16). With this chemistry, a complex is formed between the alkyne functionalized DYn-2 and the azide functionalized biotin. This biotin functional group facilitates downstream detection, enrichment, and mass spectrometry based identification (Fig. 1). In an evaluation experiment, DYn-2 was found to efficiently detect H2O2-dependent sulfenic acid modifications in recombinant glutathione peroxidase 3 (Gpx3) of budding yeast (18). Moreover, it was reported that DYn-2 is membrane permeable, non-toxic, and a non-influencer of the intracellular redox balance (17, 18). Therefore, DYn-2 has been suggested as a global sulfenome reader in living cells (17, 18), and has been applied to investigate epidermal growth factor (EGF) mediated protein sulfenylation in a human epidermoid carcinoma A431 cell line and to identify intracellular protein targets of H2O2 during cell signaling (17).Open in a separate windowFig. 1.Schematic views of the molecular mechanism of the DYn-2 probe and the strategy to identify DYn-2 trapped sulfenylated proteins. A, DYn-2 specifically detects sulfenic acid modifications, but no other thiol modifications. B, Biotinylation of the DYn-2 tagged proteins by click reaction. C, Once DYn-2 tagged proteins are biotinylated, a streptavidin-HRP (Strep-HRP) blot visualizes sulfenylation, or alternatively, after enrichment on avidin beads, proteins are identified by mass spectrometry analysis.Here, we selected the DYn-2 probe to identify the sulfenome in plant cells under oxidative stress. Through a combination of biochemical, immunoblot and mass spectrometry techniques, and TAIR10 database and SUBA3-software predictions, we can claim that DYn-2 is able to detect sulfenic acids on proteins located in different subcellular compartments of plant cells. We identified 226 sulfenylated proteins in response to an H2O2 treatment of Arabidopsis cell suspensions, of which 123 proteins are new candidates for cysteine oxidative post-translational modification (PTM) events.  相似文献   

17.
18.
Linear genome stability requires specialized telomere replication and protection mechanisms. A common solution to this problem in non-eukaryotes is the formation of hairpin telomeres by telomere resolvases (also known as protelomerases). These enzymes perform a two-step transesterification on replication intermediates to generate hairpin telomeres using an active site similar to that of tyrosine recombinases and type IB topoisomerases. Unlike phage telomere resolvases, the telomere resolvase from the Lyme disease pathogen Borrelia burgdorferi (ResT) is a permissive enzyme that resolves several types of telomere in vitro. However, the ResT region and residues mediating permissive substrate usage have not been identified. The relapsing fever Borrelia hermsii ResT exhibits a more restricted substrate usage pattern than B. burgdorferi ResT and cannot efficiently resolve a Type 2 telomere. In this study, we determined that all relapsing fever ResTs process Type 2 telomeres inefficiently. Using a library of chimeric and mutant B. hermsii/B. burgdorferi ResTs, we mapped the determinants in B. burgdorferi ResT conferring the ability to resolve multiple Type 2 telomeres. Type 2 telomere resolution was dependent on a single proline in the ResT catalytic region that was conserved in all Lyme disease but not relapsing fever ResTs and that is part of a 2-amino acid insertion absent from phage telomere resolvase sequences. The identification of a permissive substrate usage determinant explains the ability of B. burgdorferi ResT to process the 19 unique telomeres found in its segmented genome and will aid further studies on the structure and function of this essential enzyme.Replication and protection of telomeric DNA are required to ensure the genomic stability of all organisms with linear replicons. Until quite recently, it was assumed that linearity is a property confined to the replicons of eukaryotes and certain primarily eukaryotic viruses. However, a growing body of evidence indicates that linear DNA is also found in a broad range of bacteriophages (16) and in bacteria themselves (710), including the Borrelia species that cause Lyme disease and relapsing fever (11, 12). A common solution to the end replication and protection problem in non-eukaryotes is the covalent sealing of DNA ends in the form of hairpins (2, 46, 10, 11, 1316). Hairpin DNA is not recognized as a double-strand break, and continuous synthesis of DNA around the hairpin loop abolishes the end replication problem. However, mother and daughter replicons are covalently linked at the junction of their telomeres following DNA replication; separation of the two replicons and formation of new hairpin telomeres require a DNA breakage and reunion process referred to as telomere resolution (17, 18).Resolution of the linear chromosome and plasmids in Borrelia species and of the linear plasmid prophages from Escherichia coli, Yersinia enterocolitica, and Klebsiella oxytoca is performed by telomere resolvases (also referred to as protelomerases) (5, 1921). A growing number of candidate telomere resolvases have been identified in the genomes of eukaryotic viruses, phages, and bacteria (22, 23). Telomere resolvases are DNA cleavage and rejoining enzymes related to tyrosine recombinases and type 1B topoisomerases (19, 21, 22, 24, 25). Telomere resolvase catalyzes a two-step transesterification reaction in which staggered cuts are introduced 6 bp apart on either side of the axis of symmetry in the replicated telomere substrate (5, 19, 21, 24). Cleavage is accompanied by the formation of a 3′-phosphotyrosyl protein-DNA linkage. Subsequent nucleophilic attack on opposing strands by the free 5′-OH groups in the nicked substrate creates covalently closed hairpin telomeres. A recent crystal structure of the Klebsiella phage telomere resolvase (TelK) in complex with its substrate identified the residues involved in catalysis (25); all but one of these residues are conserved in all telomere resolvases (22), implying that the basic catalytic mechanism underlying telomere resolution is conserved. However, telomere resolvase sequences vary substantially outside of the central catalytic region (25, 26), and the enzymes characterized to date demonstrate important differences in substrate usage that likely reflect functionally distinct mechanisms of substrate interaction.The Borrelia burgdorferi telomere resolvase, ResT, appears to be particularly divergent. It is substantially smaller than phage telomere resolvases, and unlike its phage counterparts (5, 20, 21), it cannot efficiently resolve negatively supercoiled DNA (19, 27), presumably reflecting differences in the substrates resolved by phage and Borrelia telomere resolvases in vivo. On the other hand, B. burgdorferi ResT can fuse hairpin telomeres in a reversal of the resolution reaction (28), a function that is not shared with the phage telomere resolvase TelK (25). It can also synapse replicated telomeres and catalyze the formation of Holliday junctions (29). The ability of ResT to promote hairpin fusion has been proposed as the mechanism underlying the ongoing genetic rearrangements that are a prominent feature of the B. burgdorferi genome (18, 28). Finally, B. burgdorferi ResT can tolerate a surprising amount of variation in its substrate (30, 31), a feature that is not shared by phage telomere resolvases (21). Although B. burgdorferi ResT appears to be more permissive with a greater scope of activities than other telomere resolvases, the sequences mediating most of its unique properties have not yet been identified.The B. burgdorferi genome contains a total of 19 distinct hairpin sequences, all of which must be resolved by ResT (31). These sequences can be classified into three groups based on the presence and positioning of the box 1 motif, which is a critical determinant of activity in phage and Borrelia telomere resolvases (see Fig. 1A) (21, 24, 30). A box 1-like motif is also found in many of the hairpin telomeres sequenced to date (6, 14, 3235), although its function in telomere resolution is unknown. The box 1 consensus sequence (TAT(a/t)AT) closely resembles the −10/Pribnow box and TATA box consensus sequences of prokaryotic and eukaryotic promoters (TATAAT and TATA(a/t)A(a/t), respectively), which undergo transient deformations that predispose them to melting (36) and are intrinsically bent and anisotropically flexible (37). Therefore, box 1 may facilitate nucleation of hairpin folding and/or may confer an intrinsic bend or flexibility to substrates that is important for the resolution reaction.Open in a separate windowFIGURE 1.Species-specific resolution of Type I and 2 telomeres. A, a schematic showing the three types of hairpin telomere found on the linear replicons of the B. burgdorferi genome (see Ref. 31). The box 1 sequence in Type 1 and 2 telomeres is situated 1 and 4 nucleotides away from the axis of symmetry, respectively, whereas Type 3 telomeres contain no clear box 1. B, a schematic illustrating the telomere resolution reaction substrate and products is shown along with two ethidium bromide-stained agarose gels showing telomere resolution assays. The gels show resolution kinetics for B. burgdorferi and B. hermsii ResT on Type 1 and 2 telomeres (plasmid substrates pYT1/lp17L and pYT92/chromL, respectively).B. burgdorferi ResT can resolve telomeres in which box 1 is located at positions 1 and 4 nucleotides away from the axis of symmetry (Type 1 and 2 telomeres, respectively), as well as AT-rich telomeres without a box 1 sequence (Type 3 telomeres) (see Fig. 1A) (30, 31). B. burgdorferi ResT cleaves telomeres at a fixed position relative to the axis of symmetry, independent of the location of box 1 (30). Positioning of the enzyme for cleavage in all telomere types is most likely driven by sequence-specific interactions between ResT domains 2 (catalytic) and/or 3 (C-terminal) and a fixed element upstream of box 1 that is positioned 14 nucleotides from the axis of symmetry in all Borrelia telomeres (box 3 and adjacent nucleotides) (see Figs. 1A and and2)2) (26, 30, 31). In contrast, box 1 and axis-flanking nucleotides are not involved in high affinity and/or sequence-specific interactions with ResT and require the ResT N-terminal domain for full protection in DNase footprinting assays (26, 27). The most likely candidate for interactions with box 1 and axis-flanking nucleotides is a Borrelia-specific hairpin-binding region in the N terminus, which is thought to promote a pre-hairpinning step involving strand opening at the axis (38).Open in a separate windowFIGURE 2.Alignment of 11 Borrelia ResT sequences. Shown is ClustalW2 alignment of ResT amino acid sequences from five Lyme disease Borrelia species (B. afzelii, B. spielmanii, B. valaisiana, B. garinii, and B. burgdorferi), five relapsing fever Borrelia species (B. turicatae, B. parkeri, B. hermsii, B. recurrentis, and B. duttonii), and one avian Borrelia species (B. anserina) (generated using ClustalW2 from the EBI web site) (19, 3942, 48, 49). The sequences for B. anserina, B. parkeri, and B. turicatae ResTs are reported for the first time in this study (respective GenBank accession numbers are FJ882620, FJ882621, and FJ882623). Sequences are arranged in order of similarity to neighboring sequences and are colored in JalView using the Zappo coloring scheme for identifying amino acids with similar physicochemical properties (50). Only residues that are identical in 100% of ResTs are indicated by colored shading. Arrows above the alignment indicate ResT domain boundaries identified by chymotrypsin digest, sequence comparison with other proteins, and HHsenser predictions (26, 51). The hairpin-binding motif found in cut-and-paste transposases is indicated beneath the alignment by white text on a black background (38). The positions corresponding to the active site residues in tyrosine recombinases, type IB topoisomerases, and TelK are indicated by blue asterisks below the sequence, with the active site tyrosine nucleophile at position 335 marked by a red asterisk (22, 25). The ringed black dot below position 326 indicates an amino acid in the active site region that differs in Lyme disease and relapsing fever ResTs. Sequences above the black line drawn between B. burgdorferi and B. turicatae are from Lyme disease Borrelia species; sequences below the black line are from relapsing fever Borrelia species. The ResT sequence from the avian Borrelia species B. anserina is shown at bottom.ResT from the relapsing fever Borrelia species Borrelia hermsii exhibits a more restricted substrate usage pattern in vitro when compared with ResT from the Lyme disease pathogen B. burgdorferi (39). Specifically, B. hermsii ResT is unable to efficiently resolve a Type 2 telomere. Therefore, B. burgdorferi ResT appears to be a more permissive enzyme than its relapsing fever counterpart. In this study, we investigated the basis for permissive substrate usage by B. burgdorferi ResT. Using a library of chimeric B. hermsii/B. burgdorferi ResTs, we mapped the sequence determinants in B. burgdorferi ResT that confer the ability to resolve multiple Type 2 telomeres. Surprisingly, this approach indicated that Type 2 telomere resolution was crucially regulated by a single proline residue located in a small Borrelia-specific insertion in the central catalytic region of ResT. The proline at this position was conserved in the ResTs from all Lyme disease Borrelia species but in none of the ResTs from relapsing fever Borrelia species, which were unable to efficiently resolve Type 2 telomeres in vitro. This study has identified a specific residue in ResT responsible for permissive substrate usage patterns.  相似文献   

19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号