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.
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   

2.
3.
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   

4.
5.
Heme and chlorophyll accumulate to high levels in legume root nodules and in photosynthetic tissues, respectively, and they are both derived from the universal tetrapyrrole precursor δ-aminolevulinic acid (ALA). The first committed step in ALA and tetrapyrrole synthesis is catalyzed by glutamyl-tRNA reductase (GTR) in plants. A soybean (Glycine max) root-nodule cDNA encoding GTR was isolated by complementation of an Escherichia coli GTR-defective mutant for restoration of ALA prototrophy. Gtr mRNA was very low in uninfected roots but accumulated to high levels in root nodules. The induction of Gtr mRNA in developing nodules was subsequent to that of the gene Enod2 (early nodule) and coincided with leghemoglobin mRNA accumulation. Genomic analysis revealed two Gtr genes, Gtr1 and a 3′ portion of Gtr2, which were isolated from the soybean genome. RNase-protection analysis using probes specific to Gtr1 and Gtr2 showed that both genes were expressed, but Gtr1 mRNA accumulated to significantly higher levels. In addition, the qualitative patterns of expression of Gtr1 and Gtr2 were similar to each other and to total Gtr mRNA in leaves and nodules of mature plants and etiolated plantlets. The data indicate that Gtr1 is universal for tetrapyrrole synthesis and that a Gtr gene specific for a tissue or tetrapyrrole is unlikely. We suggest that ALA synthesis in specialized root nodules involves an altered spatial expression of genes that are otherwise induced strongly only in photosynthetic tissues of uninfected plants.Soybean (Glycine max) and numerous other legumes can establish a symbiosis with rhizobia, resulting in the formation of root nodules comprising specialized plant and bacterial cells (for review, see Mylona et al., 1995). Rhizobia reduce atmospheric nitrogen to ammonia within nodules, which is assimilated by the plant host to fulfill its nutritional nitrogen requirement. The high energy requirement for nitrogen fixation necessitates efficient respiration by the prokaryote within the microaerobic milieu of the nodule. The plant host synthesizes a nodule-specific hemoglobin (leghemoglobin) that serves to facilitate oxygen diffusion to the bacterial endosymbiont and to buffer the free oxygen concentration at a low tension (for review, see Appleby, 1992). Both of these functions require that the hemoglobin concentration be high, and, indeed, it exceeds 1 mm in soybean nodules (Appleby, 1984) and is the predominant plant protein in that organ. Once thought to be confined to legume nodules, hemoglobins are found throughout the plant kingdom, and leghemoglobin likely represents a specialization of a general plant phenomenon (for review, see Hardison, 1996). A gene encoding a nonsymbiotic hemoglobin has been identified in soybean and other legumes (Andersson et al., 1996); therefore, expression in nodules involves the specific activation of a subset of genes within a gene family. Leghemoglobin genes may have arisen from gene duplication, followed by specialization (Andersson et al., 1996).Hemes and chlorophyll are tetrapyrroles synthesized from common precursors; chlorophyll is quantitatively the major tetrapyrrole in plants, with heme and other tetrapyrroles being present in minor amounts. Legume root nodules represent an exception, in which heme is synthesized in high quantity in the absence of chlorophyll, thus requiring the activity of enzymes not normally expressed highly in nonphotosynthetic tissues. Heme is synthesized from the universal tetrapyrrole precursor ALA by seven successive enzymatic steps; chlorophyll formation diverges after the synthesis of protoporphyrin, the immediate heme precursor (for review, see O''Brian, 1996). Biochemical and genetic evidence shows that soybean heme biosynthesis genes are strongly induced in root nodules (Sangwan and O''Brian, 1991, 1992, 1993; Madsen et al., 1993; Kaczor et al., 1994; Frustaci et al., 1995; Santana et al., 1998), and immunohistochemical studies demonstrate that induction is concentrated in infected nodule cells (Santana et al., 1998).ALA is synthesized from Glu in plants by a three-step mechanism called the C5 pathway (Fig. (Fig.1);1); the latter two steps are committed to ALA synthesis and are catalyzed by GTR and GSAT, respectively (for review, see Beale and Weinstein, 1990; Jahn et al., 1991). Plant cDNA or genes encoding GTR (Gtr, also called HemA) and GSAT (Gsa) have been identified in several plant species (Grimm, 1990; Sangwan and O''Brian, 1993; Hofgen et al., 1994; Ilag et al., 1994; Frustaci et al., 1995; Wenzlau and Berry-Lowe, 1995; Bougri and Grimm, 1996; Kumar et al., 1996; Tanaka et al., 1996). Two genes for each enzyme have been described, and some genes are reported to be specific to a tissue, tetrapyrrole, or light regimen (Bougri and Grimm, 1996; Kumar et al., 1996; Tanaka et al., 1996). However, soybean Gsa1 is highly expressed in both leaves and nodules and contains a cis-acting element in its promoter that binds to a nuclear factor found in both tissues. (Frustaci et al., 1995). In this study we isolated soybean Gtr1 and characterized the genetic basis of GTR expression in root nodules. Figure 1C5 pathway for ALA synthesis. The committed steps for ALA synthesis catalyzed by GTR and GSAT are boxed. Glutamyl-tRNA synthetase (GluRS) and glutamyl-tRNAGlu also participate in protein synthesis. The gene designations in plants are shown in parentheses ...  相似文献   

6.
7.
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   

8.
9.
10.
All of the 2,6-dideoxy sugars contained within the structure of chromomycin A3 are derived from d-glucose. Enzyme assays were used to confirm the presence of hexokinase, phosphoglucomutase, UDPG pyrophosphorylase (UDPGP), and UDPG oxidoreductase (UDPGO), all of which are involved in the pathway of glucose activation and conversion into 2,6-dideoxyhexoses during chromomycin biosynthesis. Levels of the four enzymes in Streptomyces spp. cell extracts were correlated with the production of chromomycins. The pathway of sugar activation in Streptomyces spp. involves glucose 6-phosphorylation by hexokinase, isomerization to G-1-P catalyzed by phosphoglucomutase, synthesis of UDPG catalyzed by UDPGP, and formation of UDP-4-keto-6-deoxyglucose by UDPGO.Dideoxy sugars occur commonly in the structures of cardiac glycosides from plants, in antibiotics like chromomycin A3 (Fig. (Fig.1),1), and in macrolides produced by microorganisms. On the basis of stable isotope-labeling experiments, biosynthetic studies conducted in Rosazza’s laboratory have indicated that all the deoxy sugars of chromomycin A3 are derived from d-glucose (21). While the assembly of the polyketide aglycone is reasonably well understood, relatively little is known of the details of 2,6-dideoxy sugar biogenesis in streptomycetes. Earlier studies with Streptomyces rimosus indicated that TDP-mycarose is synthesized from TDP-d-glucose (TDPG) and S-adenosyl-l-methionine (10, 23). The reaction requires NADPH as a cofactor, and TDP-4-keto-6-deoxy-d-glucose is an intermediate. Formation of TDP-4-keto-6-deoxy-d-glucose was catalyzed by the enzyme TDPG oxidoreductase (TDPG-4,6-dehydratase; EC 4.2.1.46). Similar 4-keto sugar nucleotides are intermediates for the biosynthesis of polyene macrolide antibiotic amino sugars (18). Similar pathways have been elaborated for the formation of 2,6-dideoxy-d-threo-4-hexulose of granaticin in Escherichia coli (6, 25) and 2,6-dideoxy-d-arabino-hexose of chlorothricin (12). The initial 6-deoxygenation of glucose during 3,6-dideoxy sugar formation involves a similar mechanism (32). In all of these processes, glucose is first activated by conversion into a sugar nucleotide such as UDPG followed by NAD+ oxidation of the 4 position to the corresponding 4-oxo derivative. Position 6 deoxygenation involves a general tautomerization, dehydration, and NADH,H+-catalyzed reduction process (6, 12, 25). A similar tautomerization and dehydration followed by reduction may produce C-3-deoxygenated products, such as CDP-3,6-dideoxyglucose (27). The pathway for formation of 3,6-dideoxyhexoses from CDPG in Yersinia pseudotuberculosis was clearly elucidated by Liu and Thorson (14). However, none of this elegant work was focused on the earlier steps of hexose nucleotide formation. Open in a separate windowFIG. 1Structures of chromomycins A2 and A3.On the basis of previous work (7), it is reasonable to postulate that the biosynthesis of 2,6-dideoxyglucose in Streptomyces griseus involves phosphorylation to glucose-6-phosphate by hexokinase (HK; E.C.2.7.7.1), as in glycolysis; conversion to glucose-1-phosphate by phosphoglucomutase (PGM; EC 2.7.5.1); reaction with UTP to form UDPG in a reaction catalyzed by UDPG pyrophosphorylase (UDPGP) (glucose-1-phosphate uridylyltransferase; EC 2.7.7.9), and C-6 deoxygenation catalyzed by UDP-d-glucose-4,6-dehydratase with NAD+ as a cofactor (Fig. (Fig.2).2). UDPG and GDPG have been detected in cell extracts of S. griseus and Streptomyces sp. strain MRS202, suggesting that these compounds are active sugar nucleotides involved in the formation of dideoxyhexoses (15). UDPGP genes from several bacteria have been cloned and sequenced (1, 3, 4, 11, 29, 30). Although nucleotidyl diphosphohexose-4,6-dehydratases (NDP-hexose-4,6-dehydratases) have been purified and characterized from several sources (5, 8, 9, 13, 19, 25, 26, 31, 33), the occurrence of the glucose-activating enzymes HK, PGM, UDPGP, and UDPG oxidoreductase (UDPGO) involved in 2,6-dideoxyhexose formation has not been established in streptomycetes. This work provides evidence for the presence of these enzymes involved in the biosynthetic activation of glucose to the 2,6-dideoxyhexoses in chromomycin A3.Open in a separate windowFIG. 2Proposed pathway for the formation of 2,6-dideoxy sugars in streptomycetes involving HK, PGM, UDPGP, and UDPGO.  相似文献   

11.
12.
13.
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   

14.
15.
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   

16.
Long-term aging of potato (Solanum tuberosum) seed-tubers resulted in a loss of patatin (40 kD) and a cysteine-proteinase inhibitor, potato multicystatin (PMC), as well as an increase in the activities of 84-, 95-, and 125-kD proteinases. Highly active, additional proteinases (75, 90, and 100 kD) appeared in the oldest tubers. Over 90% of the total proteolytic activity in aged tubers was sensitive to trans-epoxysuccinyl-l-leucylamido (4-guanidino) butane or leupeptin, whereas pepstatin was the most effective inhibitor of proteinases in young tubers. Proteinases in aged tubers were also inhibited by crude extracts or purified PMC from young tubers, suggesting that the loss of PMC was responsible for the age-induced increase in proteinase activity. Nonenzymatic oxidation, glycation, and deamidation of proteins were enhanced by aging. Aged tubers developed “daughter” tubers that contained 3-fold more protein than “mother” tubers, with a polypeptide profile consistent with that of young tubers. Although PMC and patatin were absent from the older mother tubers, both proteins were expressed in the daughter tubers, indicating that aging did not compromise the efficacy of genes encoding PMC and patatin. Unlike the mother tubers, proteinase activity in daughter tubers was undetectable. Our results indicate that tuber aging nonenzymatically modifies proteins, which enhances their susceptibility to breakdown; we also identify a role for PMC in regulating protein turnover in potato tubers.Potato (Solanum tuberosum) seed-tubers are a model system for studying the process of aging in plants. The tubers can be stored (at 4°C and 95% RH) for to 3 years without a loss of viability. However, storage (aging) beyond about 8 months effects a progressive decline in apical dominance, rooting ability, and sprout vigor (Kumar and Knowles, 1993a). In addition to changes in growth potential, aging is accompanied by increased respiration of tubers (Kumar and Knowles, 1996a), oxidative stress (Kumar and Knowles, 1996b), lipid peroxidation (Kumar and Knowles, 1993b), and decreased protein content (Kumar and Knowles, 1993c). Although protein loss is partly due to reduced synthesis (Kumar and Knowles, 1993c), the contribution of proteolysis and the mechanisms by which proteins become damaged and subsequently targeted for degradation with advancing age are unknown. Processes that may lead to protein degradation during aging include (a) increased accessibility of proteins to proteinases resulting from decompartmentation, (b) molecular modifications to polypeptides that enhance proteolysis, and (c) increased activity of proteinases (Dalling, 1987).Oxidation, glycation, and isomerization/racemization of amino acid residues of proteins have been identified as nonenzymatic mechanisms that can adversely affect structure and function (Fig. (Fig.1),1), rendering proteins more susceptible to proteolysis during aging (Dalling, 1987; Stadtman, 1992; Luthra and Balasubramanian, 1993; Eckardt and Pell, 1995). Oxidative stress contributes to the formation of carbonyl derivatives on amino acid residues of proteins (Dalling, 1987; Oliver et al., 1987; Levine et al., 1990). For example, carbonyl content and susceptibility of Rubisco to proteolysis increased during oxidative stress (Ferriera and Shaw, 1989; Penarrubia and Moreno, 1990; Garcia-Ferris and Moreno, 1993; Eckardt and Pell, 1995). Similarly, oxidative stress caused by the inhibition of catalase by aminotriazole in maize seedlings resulted in a 2-fold increase in protein carbonyl content (Prasad, 1997). The increased oxidative stress accompanying aging of potato tubers may provide an ideal environment for oxidation of proteins. Figure 1Schematic diagram showing several nonenzymatic mechanisms that could affect protein structure and function in aging potato tubers. Protein modifications that may accompany aging include oxidation (increased carbonyl groups), glycation (reaction of ...Amino groups of proteins can react with aldehyde or keto groups of reducing sugars through a Schiff-base reaction, yielding brown fluorescent pigments known as advanced glycation end products (Luthra and Balasubramaniyan, 1993). Proteins thus modified tend to form cross-links (Fig. (Fig.1)1) that can destroy protein function (Wettlaufer and Leopold, 1991). A number of age-related diseases in humans are attributed to protein glycation. For example, in diabetics, elevated blood Glc is associated with cataracts (Monnier et al., 1979), accelerated aging, and vascular narrowing (Brownlee et al., 1986; Cerami et al., 1987). In light of the substantial increase in reducing sugar concentration of tubers during aging (Kumar and Knowles, 1993b), it was of interest to determine the extent of age-induced protein glycation.Proteins are also susceptible to nonenzymatic modification by deamidation-mediated conversion of l-asparaginyl to l-isoaspartyl residues (Fig. (Fig.1).1). Although proteins containing isomerized residues can be targeted for degradation, they are also substrates for PCMT (type II), which can restore protein function. Repair to such damaged proteins involves methylation, using AdoMet as a methyl donor. PCMT is a cytosolic “housekeeping” enzyme with specificity for the recognition and repair of altered aspartyl residues (Galletti et al., 1995), and has been detected in 45 plant species belonging to 23 families (Mudgett et al., 1997). Changes in PCMT activity with advancing tuber age were thus characterized as an indicator of deamidation-mediated damage to proteins.In addition to reduced protein synthesis and enhanced susceptibility of proteins to proteolysis, advancing tuber age may contribute to loss in the ability to synthesize proteinase inhibitors and thus to protein catabolism. Potato tubers contain a proteinase inhibitor, PMC (Rodis and Hoff, 1984; Walsh and Strikland, 1993). With its multiple inhibitory domains, PMC (85 kD) has the capacity for simultaneous inhibition of several Cys-proteinase molecules (Walsh and Strickland, 1993). The effect of aging on PMC and proteinase levels is unknown. Using potato as a model system, we examined potential mechanisms for age-induced protein loss and the extent to which proteins become nonenzymatically modified during aging.  相似文献   

17.
Suspension-cultured Chenopodium album L. cells are capable of continuous, long-term growth on a boron-deficient medium. Compared with cultures grown with boron, these cultures contained more enlarged and detached cells, had increased turbidity due to the rupture of a small number of cells, and contained cells with an increased cell wall pore size. These characteristics were reversed by the addition of boric acid (≥7 μm) to the boron-deficient cells. C. album cells grown in the presence of 100 μm boric acid entered the stationary phase when they were not subcultured, and remained viable for at least 3 weeks. The transition from the growth phase to the stationary phase was accompanied by a decrease in the wall pore size. Cells grown without boric acid or with 7 μm boric acid were not able to reduce their wall pore size at the transition to the stationary phase. These cells could not be kept viable in the stationary phase, because they continued to expand and died as a result of wall rupture. The addition of 100 μm boric acid prevented wall rupture and the wall pore size was reduced to normal values. We conclude that boron is required to maintain the normal pore structure of the wall matrix and to mechanically stabilize the wall at growth termination.The ultrastructure and physical properties of plant cell walls are known to be affected by boron deficiency (Kouchi and Kumazawa, 1976; Hirsch and Torrey, 1980; Fischer and Hecht-Buchholz, 1985; Matoh et al., 1992; Hu and Brown, 1994; Findeklee and Goldbach, 1996). Moreover, boron is predominantly localized in the cell wall when plants are grown with suboptimal boron (Loomis and Durst, 1991; Matoh et al., 1992; Hu and Brown, 1994; Hu et al., 1996). In radish, >80% of the cell wall boron is present in the pectic polysaccharide RG-II (Matoh et al., 1993; Kobayashi et al., 1996), which is now known to exist as a dimer that is cross-linked by a borate ester between two apiosyl residues (Kobayashi et al., 1996; O''Neill et al., 1996). Dimeric RG-II is unusually stable at low pH and is present in a large number of plant species (Ishii and Matsunaga, 1996; Kobayashi et al., 1996, 1997; Matoh et al., 1996; O''Neill et al., 1996; Pellerin et al., 1996; Kaneko et al., 1997). The widespread occurrence and conserved structure of RG-II (Darvill et al., 1978; O''Neill et al., 1990) have led to the suggestion that borate ester cross-linked RG-II is required for the development of a normal cell wall (O''Neill et al., 1996; Matoh, 1997).One approach for determining the function of boron in plant cell walls is to compare the responses to boron deficiency of growing plant cells that are dividing and synthesizing primary cell walls with those of growth-limited plant cells in which the synthesis of primary cell walls is negligible. Suspension-cultured cells are well suited for this purpose because they may be reversibly transferred from a growth phase to a stationary phase. Continuous cell growth phase is maintained by frequent transfer of the cells into new growth medium (King, 1981; Kandarakov et al., 1994), whereas a stationary cell population is obtained by feeding the cells with Suc and by not subculturing them. Cells in the stationary phase are characterized by mechanically stabilized primary walls and reduced biosynthetic activity. Here we describe the responses of suspension-cultured Chenopodium album L. cells in the growth and stationary phases to boron deficiency. These cells have a high specific-growth rate, no significant lag phase, and reproducible changes in their wall pore size during the transition from the growth phase to the stationary phase (Titel et al., 1997).  相似文献   

18.
19.
Folding and stability are parameters that control protein behavior. The possibility of conferring additional stability on proteins has implications for their use in vivo and for their structural analysis in the laboratory. Cyclic polypeptides ranging in size from 14 to 78 amino acids occur naturally and often show enhanced resistance toward denaturation and proteolysis when compared with their linear counterparts. Native chemical ligation and intein-based methods allow production of circular derivatives of larger proteins, resulting in improved stability and refolding properties. Here we show that circular proteins can be made reversibly with excellent efficiency by means of a sortase-catalyzed cyclization reaction, requiring only minimal modification of the protein to be circularized.Sortases are bacterial enzymes that predominantly catalyze the attachment of surface proteins to the bacterial cell wall (1, 2). Other sortases polymerize pilin subunits for the construction of the covalently stabilized and covalently anchored pilus of the Gram-positive bacterium (35). The reaction catalyzed by sortase involves the recognition of short 5-residue sequence motifs, which are cleaved by the enzyme with the concomitant formation of an acyl enzyme intermediate between the active site cysteine of sortase and the carboxylate at the newly generated C terminus of the substrate (1, 68). In many bacteria, this covalent intermediate can be resolved by nucleophilic attack from the pentaglycine side chain in a peptidoglycan precursor, resulting in the formation of an amide bond between the pentaglycine side chain and the carboxylate at the cleavage site in the substrate (9, 10). In pilus construction, alternative nucleophiles such as lysine residues or diaminopimelic acid participate in the transpeptidation reaction (3, 4).When appended near the C terminus of proteins that are not natural sortase substrates, the recognition sequence of Staphylococcus aureus sortase A (LPXTG) can be used to effectuate a sortase-catalyzed transpeptidation reaction using a diverse array of artificial glycine-based nucleophiles (Fig. 1). The result is efficient installation of a diverse set of moieties, including lipids (11), carbohydrates (12), peptide nucleic acids (13), biotin (14), fluorophores (14, 15), polymers (16), solid supports (1618), or peptides (15, 19) at the C terminus of the protein substrate. During the course of our studies to further expand sortase-based protein engineering, we were struck by the frequency and relative ease with which intramolecular transpeptidation reactions were occurring. Specifically, proteins equipped with not only the LPXTG motif but also N-terminal glycine residues yielded covalently closed circular polypeptides (Fig. 1). Similar reactivity using sortase has been described in two previous cases; however, rigorous characterization of the circular polypeptides was absent (16, 20). The circular proteins in these reports were observed as minor components of more complex reaction mixtures, and the cyclization reaction itself was not optimized.Open in a separate windowFIGURE 1.Protein substrates equipped with a sortase A recognition sequence (LPXTG) can participate in intermolecular transpeptidation with synthetic oligoglycine nucleophiles (left) or intramolecular transpeptidation if an N-terminal glycine residue is present (right).Here we describe our efforts toward applying sortase-catalyzed transpeptidation to the synthesis of circular and oligomeric proteins. This method has general applicability, as illustrated by successful intramolecular reactions with three structurally unrelated proteins. In addition to circularization of individual protein units, the multiprotein complex AAA-ATPase p97/VCP/CDC48, with six identical subunits containing the LPXTG motif and an N-terminal glycine, was found to preferentially react in daisy chain fashion to yield linear protein fusions. The reaction exploited here shows remarkable similarities to the mechanisms proposed for circularization of cyclotides, small circular proteins that have been isolated from plants (2123).  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