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.
Cystic Fibrosis (CF) is an autosomal recessive disorder caused by mutations in the gene encoding the Cystic fibrosis transmembrane conductance regulator (CFTR). ΔF508-CFTR, the most common disease-causing CF mutant, exhibits folding and trafficking defects and is retained in the endoplasmic reticulum, where it is targeted for proteasomal degradation. To identify signaling pathways involved in ΔF508-CFTR rescue, we screened a library of endoribonuclease-prepared short interfering RNAs (esiRNAs) that target ∼750 different kinases and associated signaling proteins. We identified 20 novel suppressors of ΔF508-CFTR maturation, including the FGFR1. These were subsequently validated by measuring channel activity by the YFP halide-sensitive assay following shRNA-mediated knockdown, immunoblotting for the mature (band C) ΔF508-CFTR and measuring the amount of surface ΔF508-CFTR by ELISA. The role of FGFR signaling on ΔF508-CFTR trafficking was further elucidated by knocking down FGFRs and their downstream signaling proteins: Erk1/2, Akt, PLCγ-1, and FRS2. Interestingly, inhibition of FGFR1 with SU5402 administered to intestinal organoids (mini-guts) generated from the ileum of ΔF508-CFTR homozygous mice resulted in a robust ΔF508-CFTR rescue. Moreover, combination of SU5402 and VX-809 treatments in cells led to an additive enhancement of ΔF508-CFTR rescue, suggesting these compounds operate by different mechanisms. Chaperone array analysis on human bronchial epithelial cells harvested from ΔF508/ΔF508-CFTR transplant patients treated with SU5402 identified altered expression of several chaperones, an effect validated by their overexpression or knockdown experiments. We propose that FGFR signaling regulates specific chaperones that control ΔF508-CFTR maturation, and suggest that FGFRs may serve as important targets for therapeutic intervention for the treatment of CF.Cystic fibrosis (CF)1 is a pleiotropic disease caused by an abnormal ion transport in the secretory epithelia lining the tubular organs of the body such as lungs, intestines, pancreas, liver, and male reproductive tract. In the airways of CF patients, reduced Cl and bicarbonate secretion caused by lack of functional Cystic fibrosis transmembrane conductance regulator (CFTR) on the apical surface, and hyper-absorption of Na+ because of elevated activity of ENaC (1), lead to a dehydration of the airway surface liquid (ASL). This reduces the viscosity of the mucus layer and the deposited layer of thickened mucus creates an environment that promotes bacterial colonization, which eventually leads to chronic infection of the lungs and death (2, 3).CFTR is a transmembrane protein that functions as a cAMP-regulated, ATP-dependent Cl channel that also allows passage of bicarbonate through its pore (4, 5). It also possesses ATPase activity important for Cl conductance (6, 7). The CFTR structure is predicted to consist of five domains: two membrane spanning domains (MSD1, MSD2), each composed of six putative transmembrane helices, two nucleotide binding domains (NBD1, NBD2), and a unique regulatory (R) region (8).More than 1900 CFTR mutations have been identified to date (www.genet.sickkids.on.ca/cftr). The most common mutation is a deletion of phenylalanine at position 508 (ΔF508 or ΔF508-CFTR) in NBD1 (9). The ΔF508 mutation causes severe defects in the processing and function of CFTR. The protein exhibits impaired trafficking from the endoplasmic reticulum (ER) to the plasma membrane (PM), impaired intramolecular interactions between NBD1 and the transmembrane domain, and cell surface instability (1015). Nevertheless, the ΔF508 defect can be corrected, because treating cells expressing ΔF508-CFTR with low temperature or chemical chaperones (e.g. glycerol) can restore some surface expression of the mutant (11, 16).Numerous small molecules that can at least partially correct (or potentiate) the ΔF508-CFTR defect have been identified to date (1727), and some were already tested in clinical trials (e.g. sildenafil, VX-809/Lumacaftor), or have made it to the clinic (VX-770/Kalydeco/Ivacaftor) (http://www.cff.org/research/DrugDevelopmentPipeline/). However, the need to identify new ΔF508-CFTR correctors remains immense as the most promising corrector, VX-809, has proven ineffective in alleviating lung disease of CF patients when administered alone (27). Thus, our group developed a high-content technology aimed at identifying proteins and small molecules that correct the trafficking and functional defects of ΔF508-CFTR (28). We successfully used this approach to carry out three separate high-content screens: a protein overexpression screen (28), a small-molecule kinase inhibitor screen (29) and a kinome RNA interference (RNAi) screen, described here.  相似文献   

4.
Rev3 polymerase and Mph1 DNA helicase participate in error-prone and error-free pathways, respectively, for the bypassing of template lesions during DNA replication. Here we have investigated the role of these pathways and their genetic interaction with recombination factors, other nonreplicative DNA helicases, and DNA damage checkpoint components in the maintenance of genome stability, viability, and sensitivity to the DNA-damaging agent methyl methanesulfonate (MMS). We find that cells lacking Rev3 and Mph1 exhibit a synergistic, Srs2-dependent increase in the rate of accumulating spontaneous, gross chromosomal rearrangements, suggesting that the suppression of point mutations by deletion of REV3 may lead to chromosomal rearrangements. While mph1Δ is epistatic to homologous recombination (HR) genes, both Rad51 and Rad52, but not Rad59, are required for normal growth of the rev3Δ mutant and are essential for survival of rev3Δ cells during exposure to MMS, indicating that Mph1 acts in a Rad51-dependent, Rad59-independent subpathway of HR-mediated lesion bypass. Deletion of MPH1 helicase leads to synergistic DNA damage sensitivity increases in cells with chl1Δ or rrm3Δ helicase mutations, whereas mph1Δ is hypostatic to sgs1Δ. Previously reported slow growth of mph1Δ srs2Δ cells is accompanied by G2/M arrest and fully suppressed by disruption of the Mec3-dependent DNA damage checkpoint. We propose a model for replication fork rescue mediated by translesion DNA synthesis and homologous recombination that integrates the role of Mph1 in unwinding D loops and its genetic interaction with Rev3 and Srs2-regulated pathways in the suppression of spontaneous genome rearrangements and in mutation avoidance.Nonreplicative DNA helicases play an important role in the maintenance of genome stability from bacteria to humans, most likely by affecting the formation and/or resolution of recombination intermediates and by facilitating replication fork progression through chromosomal regions with a propensity to adopt unusual DNA structures or those bound by proteins. In Saccharomyces cerevisiae, this group of DNA helicases includes the 3′-to-5′ helicases Sgs1 and Srs2 and the 5′-to-3′ DNA helicase Rrm3. In the absence of any two of these three helicases, unresolved recombination intermediates accumulate and lead to extremely slow growth that is fully suppressed by deletion of genes encoding early homologous recombination (HR) factors (4, 6, 17, 20, 37, 46). In the absence of Sgs1, cells exhibit increased rates of mitotic recombination, frequent chromosome missegregation, accumulation of extrachromosomal ribosomal DNA (rDNA) circles, and increased rates of gross chromosomal rearrangements (GCRs) involving nonhomologous chromosomes (5, 24, 25, 38, 40, 43, 49, 50). Based on the increased crossover frequency during HO endonuclease-induced double-strand breaks (DSBs) in cells lacking Sgs1, it has also been proposed that Sgs1 may function in decatenation of Holliday junctions (HJs) to yield noncrossovers (12, 22). Like Sgs1, Srs2 acts to favor noncrossover outcomes during DSB repair but appears to act earlier than Sgs1 in regulating recombination outcomes through its ability to dislodge Rad51 from recombinogenic 3′ overhangs, thereby promoting a noncrossover synthesis-dependent single-strand annealing (SDSA) pathway (12, 33, 35). In contrast, Rrm3 has not been implicated in DNA repair but is thought to be important for avoidance of recombination substrate formation by removal of DNA protein complexes in certain chromosomal locations, such as chromosome ends and replication fork barriers at the rDNA locus, thus facilitating replication fork progression (13, 14).In addition to Sgs1, Rrm3, and Srs2, the yeast genome encodes two other nonreplicative DNA helicases with proposed functions in DNA repair, Mph1 and Chl1. Mph1 possesses 3′-to-5′ helicase activity, and its ATPase activity requires a relatively long fragment of single-stranded DNA (ssDNA) (≥40 nucleotides [nt]) for full activity in vitro (32). Mph1 is also necessary for resistance to the DNA damaging agents methyl methanesulfonate (MMS) and 4-nitroquinoline-1-oxide (4-NQO) and suppresses spontaneous mutations toward canavanine resistance (3, 41). The modest mutator phenotype of the mph1Δ mutant is enhanced by additional mutations in base excision repair (apn1Δ and apn2Δ) and is suppressed by mutations in translesion DNA synthesis (TLS) (rev3Δ) (36, 41). These findings, in combination with the observation of an epistatic relationship between mph1Δ and homologous recombination mutations, have led to the proposal that Mph1 may act in Rad52-dependent, error-free bypassing of DNA lesions (41). Like the 3′-to-5′ DNA helicases Sgs1 and Srs2, Mph1 was recently shown to affect crossover frequency during repair of an HO endonuclease-induced DNA DSB, favoring noncrossovers as the outcome (33). The authors showed that Mph1 can unwind intermediates of homologous recombination in vitro, specifically D loops that are thought to form early during homologous recombination when a homoduplex is invaded by a Rad51 filament. While Srs2 has been shown to be able to disassemble Rad51 filaments in vitro, it does not appear to possess Mph1''s ability to dissociate D loops once they have formed (19, 47).Although Chl1 has been shown to be required for the establishment of sister chromatid cohesion, a possible role in DNA repair by homologous recombination has also been proposed (11, 28, 30, 42). While Chl1 possesses a conserved helicase domain, helicase activity has so far been shown only for its putative human homolog, hCHLR1 (10).To further elucidate the functional interaction between nonreplicative DNA helicases and DNA repair pathways, we generated a series of mutants with combinations of mph1Δ, chl1Δ, rrm3Δ, srs2Δ, and sgs1Δ mutations and mutations in translesion DNA synthesis (TLS), base excision repair (BER), homologous recombination (HR), and DNA damage checkpoints. In addition to synthetic fitness defects due to aberrant HR and checkpoint activation, we identified epistatic and synergistic relationships with regard to fitness, the accumulation of gross chromosomal rearrangements (GCRs), and sensitivity to DNA damage. We propose that Mph1 functions in a Rad51-dependent, Rad59-independent pathway of HR for DNA lesion bypass and interacts genetically with REV3 in the suppression of gross chromosomal rearrangements.  相似文献   

5.
6.
7.
8.
NAD+ (nicotinamide adenine dinucleotide) is an essential cofactor involved in various biological processes including calorie restriction-mediated life span extension. Administration of nicotinamide riboside (NmR) has been shown to ameliorate deficiencies related to aberrant NAD+ metabolism in both yeast and mammalian cells. However, the biological role of endogenous NmR remains unclear. Here we demonstrate that salvaging endogenous NmR is an integral part of NAD+ metabolism. A balanced NmR salvage cycle is essential for calorie restriction-induced life span extension and stress resistance in yeast. Our results also suggest that partitioning of the pyridine nucleotide flux between the classical salvage cycle and the NmR salvage branch might be modulated by the NAD+-dependent Sir2 deacetylase. Furthermore, two novel deamidation steps leading to nicotinic acid mononucleotide and nicotinic acid riboside production are also uncovered that further underscore the complexity and flexibility of NAD+ metabolism. In addition, utilization of extracellular nicotinamide mononucleotide requires prior conversion to NmR mediated by a periplasmic phosphatase Pho5. Conversion to NmR may thus represent a strategy for the transport and assimilation of large nonpermeable NAD+ precursors. Together, our studies provide a molecular basis for how NAD+ homeostasis factors confer metabolic flexibility.The pyridine nucleotide NAD+ and its reduced form NADH are primary redox carriers involved in metabolism. In addition to serving as a coenzyme in redox reactions, NAD+ also acts as a cosubstrate in protein modification reactions including deacetylation and ADP-ribosylation (1, 2). NAD+ also plays an important role in calorie restriction (CR)2-mediated life span extension via regulating NAD+-dependent longevity factors (3, 4). CR is the most effective regimen known to extend life span in various species (5, 6). CR also ameliorates many age-related diseases such as cancer and diabetes (5). The Sir2 family proteins are NAD+-dependent protein deacetylases, which have been shown to play important roles in several CR models in yeast (3, 7) and higher eukaryotes (8, 9). By coupling the cleavage of NAD+ and deacetylation of target proteins, the Sir2 family proteins serve as a molecular link relaying the cellular energy state to the machinery of life span regulation. Mammalian Sir2 family proteins (SIRT1–7) have also been implicated in stress response, cell survival, and insulin and fat metabolism (810), supporting a role for SIRT proteins in age-related metabolic diseases and perhaps human aging.In eukaryotes, NAD+ is generated by de novo synthesis and by salvaging various intermediary precursors (see Fig. 1A). In yeast, the de novo pathway is mediated by Bna1–5 and Qpt1 (Bna6), which produces nicotinic acid mononucleotide (NaMN) from tryptophan (11). Because the de novo pathway requires molecular oxygen as a substrate, cells grown under anaerobic growth conditions would rely on exogenous NAD+ precursors for the nicotinamide (Nam) moiety (11). Yeast cells also salvage Nam from NAD+ consuming reactions or nicotinic acid (NA) from environment via Tna1, Pnc1, and Npt1, leading to NaMN production. NaMN is then converted to NAD+ via Nma1/2 and Qns1 (see Fig. 1A). Nma1/2 are adenylyltransferases with dual specificity toward NMN and NaMN (12, 13), and Qns1 is a glutamine-dependent NAD+ synthetase. Recent studies also showed that supplementing nicotinamide riboside (NmR) and nicotinic acid riboside (NaR) to growth medium rescued the lethality of NAD+ auxotrophic mutants (1416). Assimilations of exogenous NmR and NaR are mainly mediated by a conserved NmR kinase (Nrk1) and three nucleosidases (Urh1, Pnp1, and Meu1). Nrk1 phosphorylates NmR and NaR to produce nicotinamide mononucleotide (NMN) and NaMN, respectively (14, 16). Urh1, Pnp1, and Meu1 catabolize NmR and NaR to generate Nam and NA (15, 16).Open in a separate windowFIGURE 1.Nicotinamide riboside (NmR) is an endogenous metabolite in yeast. A, the current model of the NAD+ biosynthesis pathways. Extracellular NmR enters the salvage cycle through Nrk1, Urh1, Pnp1, and Meu1. B, NAD+ prototrophic cells release metabolites into growth medium to cross-feed NAD+ auxotrophic cells (the npt1Δqpt1Δ and qns1Δ mutants). Micro-colonies of the NAD+ auxotrophic mutants become visible after 2-day incubation at 30 °C, which show “gradient” growth patterns descending from the side adjacent to WT. C, Nrk1 is required for NAD+ auxotrophic cells to utilize NmR. Anaerobic growth conditions (−O2) are utilized to block de novo NAD+ biosynthesis in the npt1Δ and npt1Δnrk1Δ mutants. D, Nrk1 is required to utilize cross-feeding metabolites. E, cross-feeding activity is modulated by factors in NmR metabolism. Cells defective in NmR utilization (left panel) or transport (middle panel) show increased cross-feeding in spot assays. Overexpressing Nrk1 decreases cross-feeding activity (right panel). The results show growth of the npt1Δqpt1Δ recipient (plated on YPD at a density of ∼9000 cells/cm2) supported by feeder cells (∼2 × 104 cells spotted directly onto the recipient lawn). oe, overexpression.NmR supplementation has recently been shown to be a promising strategy for prevention and treatment of certain diseases (17). For example, NmR protected neurons from axonal degeneration via functioning as a NAD+ precursor (18, 19). Given that several NmR assimilating enzymes and NmR transporters have been characterized and many are conserved from fungi to mammals (14, 15, 2022), NmR has been speculated to be an endogenous NAD+ precursor (17, 23). Here, we provided direct evidence for endogenous NmR as an integral part of NAD+ metabolism in yeast. We also determined the biological significance of salvaging endogenous NmR and studied its role in CR-induced life span extension. Moreover, we demonstrated that the NmR salvage machinery was also required for utilizing exogenous NMN, which has recently been shown to increase NAD+ levels in mammalian cells (24). Finally, we discussed the role of Sir2 in modulating the flux of pyridine nucleotides between alternate routes.  相似文献   

9.
10.
11.
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   

12.
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   

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

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

16.
The N-linked glycosylation of the constant fragment (Fc) of immunoglobulin G has been shown to change during pathological and physiological events and to strongly influence antibody inflammatory properties. In contrast, little is known about Fab-linked N-glycosylation, carried by ∼20% of IgG. Here we present a high-throughput workflow to analyze Fab and Fc glycosylation of polyclonal IgG purified from 5 μl of serum. We were able to detect and quantify 37 different N-glycans by means of MALDI-TOF-MS analysis in reflectron positive mode using a novel linkage-specific derivatization of sialic acid. This method was applied to 174 samples of a pregnancy cohort to reveal Fab glycosylation features and their change with pregnancy. Data analysis revealed marked differences between Fab and Fc glycosylation, especially in the levels of galactosylation and sialylation, incidence of bisecting GlcNAc, and presence of high mannose structures, which were all higher in the Fab portion than the Fc, whereas Fc showed higher levels of fucosylation. Additionally, we observed several changes during pregnancy and after delivery. Fab N-glycan sialylation was increased and bisection was decreased relative to postpartum time points, and nearly complete galactosylation of Fab glycans was observed throughout. Fc glycosylation changes were similar to results described before, with increased galactosylation and sialylation and decreased bisection during pregnancy. We expect that the parallel analysis of IgG Fab and Fc, as set up in this paper, will be important for unraveling roles of these glycans in (auto)immunity, which may be mediated via recognition by human lectins or modulation of antigen binding.Immunoglobulins are key players of the human immune system. Immunoglobulin G (IgG)1 is the most abundant representative of this group, with serum concentrations of ∼10 mg/ml (1). It consists of two heavy chains (γ-chains) made up of three constant regions (CH1, CH2, and CH3) and one variable region (VH). Attached to each heavy chain is a light chain (λ or κ). Based on chemical and biological properties, different regions can be distinguished in the IgG molecule: two antigen binding fragments (obtained as F(ab′)2 by IdeS treatment; herein referred to as Fab) and a crystallizable fragment (Fc). The structure of IgG is schematically presented in Fig. 1.Open in a separate windowFig. 1.Schematic representation of IgG with the heavy γ chains (dark blue), light chains (lighter blue), and N-glycans. In the top right-hand corner of the Fc and Fab areas, the percentages of galactosylation, sialylation, bisection, and fucosylation are depicted. The inset represents the stable heptasaccharide core with possible extensions.IgGs are glycoproteins, and N-glycans are present at Asn297 of the CH2 domain. These glycans consist of a constant heptasaccharide core that is often modified by a core fucose and is in part decorated with bisecting N-acetylglucosamine (GlcNAc), galactose(s), and sialic acid(s) (Fig. 1) (1). The Fc glycans have been extensively studied, and glycosylation changes have been found to be associated with disease (e.g. rheumatoid arthritis) (2, 3) and aging (46). Several immune regulatory properties have been demonstrated for IgG Fc glycans (713). For example, Fc-linked glycans influence the IgG effector function by altering the three-dimensional structure of the protein, and thereby the binding to Fcγ-receptors (12, 13). Additionally, glycan–glycan interactions occur between IgG and Fcγ-receptor-IIIa (8), with the presence of a core fucose decreasing this affinity by ∼2 orders of magnitude (7).The Fab portion consists of the heavy chain CH1 and VH regions combined with a light chain and exhibits the antigen binding sites formed by the variable and hypervariable regions of those two chains. N-glycans are known to occur on 15% to 25% of the IgG Fab portions (1, 14, 15). The Fab N-glycans can be involved in immunomodulation, because they influence the affinity and avidity of antibodies for antigens (1619), as well as antibody half-life (17, 20). The glycans of the Fab have been described as biantennary complex-type structures that are, in contrast to Fc glycans, highly sialylated (2123). Additionally, high-mannose-type structures have been said to be located on the Fab portion (23).Pregnancy is known to be associated with overall changes in IgG glycosylation. Indeed, a marked increase of galactosylation and sialylation has been observed in IgG Fc glycosylation during pregnancy (3, 24, 25). In addition, lectin binding studies suggest changes in Fab glycosylation of IgG during pregnancy (26), which may be caused by increased levels of progesterone (27). Changes in glycosylation during pregnancy could be one of the mechanisms that contribute to acceptance of the fetal allograft by the maternal immune system (26).Our knowledge on the Fab glycosylation of IgGs from peripheral blood is scarce, which is in part due to difficulty detecting the glycans in a Fab-region-specific manner. Because of the polyclonal nature of serum IgG, one may expect Fab glycans to be attached to a large variety of sequence motifs arising from somatic rearrangements and mutations (28), making the analysis of Fab glycopeptides from polyclonal serum IgG very demanding, if feasible at all. Therefore, study of the Fab glycosylation of polyclonal serum IgG has mainly been pursued at the level of released glycans (14, 23). Difficulties lie in the purification of IgG and the separation of Fc and Fab glycosylation, which is essential for the assignment of the glycans to either part of the IgG molecule.Here we present a high-throughput method for studying Fab glycosylation at the level of released glycans obtained from serum-derived polyclonal IgG. Using state-of-the-art affinity capturing beads and enzymes, we were able to obtain Fab and Fc separately, which, after glycan release, resulted in Fc- and Fab-specific glycan pools. The released glycans were subjected to a novel derivatization protocol resulting in linkage-specific modification of sialic acids, followed by HILIC sample purification and MALDI-TOF-MS. Finally, because marked changes in glycosylation during pregnancy have been described, the technique was applied to consecutive serum samples from a cohort of pregnant women. This approach was chosen to determine the usefulness of this technique in a clinical setting. The method proved to be able to demonstrate pregnancy-related changes in glycosylation of the Fab portion, in addition to the already known changes in Fc glycosylation (3, 24, 25).  相似文献   

17.
18.
N-terminal acetylation (Nt-acetylation) occurs on the majority of eukaryotic proteins and is catalyzed by N-terminal acetyltransferases (NATs). Nt-acetylation is increasingly recognized as a vital modification with functional implications ranging from protein degradation to protein localization. Although early genetic studies in yeast demonstrated that NAT-deletion strains displayed a variety of phenotypes, only recently, the first human genetic disorder caused by a mutation in a NAT gene was reported; boys diagnosed with the X-linked Ogden syndrome harbor a p.Ser37Pro (S37P) mutation in the gene encoding Naa10, the catalytic subunit of the NatA complex, and suffer from global developmental delays and lethality during infancy. Here, we describe a Saccharomyces cerevisiae model developed by introducing the human wild-type or mutant NatA complex into yeast lacking NatA (NatA-Δ). The wild-type human NatA complex phenotypically complemented the NatA-Δ strain, whereas only a partial rescue was observed for the Ogden mutant NatA complex suggesting that hNaa10 S37P is only partially functional in vivo. Immunoprecipitation experiments revealed a reduced subunit complexation for the mutant hNatA S37P next to a reduced in vitro catalytic activity. We performed quantitative Nt-acetylome analyses on a control yeast strain (yNatA), a yeast NatA deletion strain (yNatA-Δ), a yeast NatA deletion strain expressing wild-type human NatA (hNatA), and a yeast NatA deletion strain expressing mutant human NatA (hNatA S37P). Interestingly, a generally reduced degree of Nt-acetylation was observed among a large group of NatA substrates in the yeast expressing mutant hNatA as compared with yeast expressing wild-type hNatA. Combined, these data provide strong support for the functional impairment of hNaa10 S37P in vivo and suggest that reduced Nt-acetylation of one or more target substrates contributes to the pathogenesis of the Ogden syndrome. Comparative analysis between human and yeast NatA also provided new insights into the co-evolution of the NatA complexes and their substrates. For instance, (Met-)Ala- N termini are more prevalent in the human proteome as compared with the yeast proteome, and hNatA displays a preference toward these N termini as compared with yNatA.Up to 85% of soluble eukaryotic proteins carry an N-terminal acetyl group at their N terminus, which is the result of a co-translational protein modification referred to as N-terminal protein acetylation (Nt-acetylation) or Nα-acetylation (1). This presumed irreversible protein modification is catalyzed by a specific category of the GCN5-related N-acetyltransferase domain containing superfamily of acetyltransferases; the ribosome associated N-terminal acetyltransferases or NATs1 (2). NATs catalyze the acetyl transfer from acetyl coenzyme A (Ac-CoA) to a primary α-amine of the first amino acid residue of a nascent protein chain. In eukaryotes, NATs are composed of at least one catalytic subunit and mainly target different substrate N termini based on their N-terminal sequences (3).To date, five human NATs hNatA, hNatB, and hNatC; constituting the major human NAT complexes, and hNatD and hNatF have been identified and their substrate specificity characterized (1, 48). In addition, a putative hNatE complex has been described (910). Except for NatF, which is only expressed in higher eukaryotes (1), the substrate specificity profiles of the NatA-E complexes seem to be conserved among eukaryotes (59, 1113).Contrary to the original assumption that Nt-acetylation protected proteins from degradation (14), it was more recently demonstrated that this modification creates specific degradation signals (termed Ac/N-degrons) in cellular proteins, thereby diversifying this original view substantially. These degrons target at least some Nt-acetylated proteins for the conditional degradation by a novel branch of the N-end rule pathway, an ubiquitin-dependent proteolytic system (1516). In addition, numerous reports implicate Nt-acetylation in cellular differentiation, survival, metabolism, and proliferation, thereby linking it to cancer (1718). As such, Nt-acetylation is now linked to a whole range of molecular implications including protein destabilization and degradation by the Nt-acetylation dependent recruitment of ubiquitin ligases (1516), protein translocation (19), membrane attachment (20), and protein complex formation (21).Among all characterized NATs, NatA displays the broadest substrate specificity profile and thus represents the primary NAT in terms of substrate N termini as it is responsible for the Nt-acetylation of the methionine aminopeptidase (MetAP) iMet-processed serine, threonine, alanine, glycine, and valine starting N termini (3). The human NatA complex is composed of two essential subunits; the catalytic subunit hNaa10 (hARD1) and the regulatory subunit hNaa15 (NATH/hNAT1) (4). Deregulations of hNaa10 and/or NatA expression have been linked to various signaling molecules including hypoxia inducible factor-1α, DNA methyltransferase1/E-cadherin, β-catenin/cyclin D1, and Bcl-xL, showing its involvement in hypoxia, tumorigenesis, cell cycle progression, and apoptosis (17, 2226).Recently, the first structures of NATs and a NAT-complex were solved, providing a molecular understanding of the sequence specific Nt-acetylation of protein N termini (2730). Structural analyses of noncomplexed Naa10 and NatA from Schizosaccharomyces pombe reveal an allosteric modulator function of Naa15 in steering Naa10 specificity and provide a rational for the distinctive substrate specificity profiles observed when assaying non-complexed versus complexed Naa10 (10, 27), with both forms co-existing in cells (10). In particular, three essential catalytic Naa10 residues were found to be incorrectly positioned in non-complexed Naa10, while these shift into the active site in Naa15-complexed Naa10, thereby permitting canonical NatA-mediated Nt-acetylation. Interestingly, noncomplexed Naa10 was shown to efficiently Nt-acetylate glutamate and aspartate starting N termini, whereas poorly acetylating canonical NatA type N termini (10). The study of Liszczak et al. further showed that NatA substrate binding specificity was coupled to the catalytic mechanism being used (27). More specifically, an essential glutamate residue (Glu24 in the protein accession Q9UTI3 (Swiss-Prot)) involved in catalysis, precludes methionine from entering the specificity pocket, whereas cognate NatA substrate N-terminal residues can easily be accommodated. Interestingly, and in contrast to NatA, both wild-type Naa10 and Glu24 mutated Naa10 (Naa10 E24A) were still capable of Nt-acetylating acidic amino acid starting N termini, most likely because of the substrate side-chain carboxyl moiety acting as a functional replacement group in the process of catalysis, whereas essentially no activity could be observed when probing a cognate NatA substrate (27).Early yeast studies demonstrated that strains with mutated or deleted NAT genes were viable, but displayed a number of different phenotypes (31). For NatA, the first phenotypes described were defects in sporulation, mating, and entry into stationary phase when NAA10 (ARD1) was mutated (32). Four years later, the overlapping phenotypes of NAA10 and NAA15 (NAT1) mutant strains, revealed, along with other data, that Naa10 and Naa15 are in fact components of the NatA acetyltransferase complex (3334). As compared with NatA phenotypes, NatB phenotypes are more severe, including slow growth and defects in mitochondrial inheritance (3536). NatC subunits were initially found to be essential for propagation of the l-A dsRNA virus, and further for growth on nonfermentable carbon sources (3739). The first reports implicating NAT gene point mutations in human genetic disorders only recently emerged. More specifically, two different point mutations in the X-linked NAA10 gene were both found to cause developmental delays and were linked to the Ogden syndrome (S37P) (40) and intellectual disability (R116W) (41), highlighting the essential importance of NATs and protein Nt-acetylation in biology and disease. Further, in Caenorhabditis elegans (42), Drosophila melanogaster (43), and Trypanosoma brucei (44), Naa10 was proven to be essential and, strengthened by the observed detrimental effects of NAA10 mutations (4041), the NAA10 gene function is also believed to be essential in human.Ogden syndrome boys harboring the p.Ser37Pro variant in the gene encoding Naa10 are characterized by craniofacial abnormalities, failure to thrive, developmental delay, hypotonia, cardiac arrhythmias, cryptorchidism, and an aged appearance, ultimately resulting in mortality during infancy (40). Although this mutation was shown to significantly impair Naa10 catalytic activity in vitro, we here assessed the influence and functional in vitro and in vivo consequences of this mutation on NatA complex formation and NatA activity in a yeast model. By phenotypic screening in yeast, we show that hNaa10 S37P displays a significantly impaired functionality in vivo. Further, using immunoprecipitation, we show that the human Naa10-Naa15 complex formation is negatively affected by the S37P mutation, and that immunoprecipitated hNatA S37P also displays a reduced in vitro catalytic activity as compared with wild-type hNatA. Finally, quantitative Nt-acetylome analyses suggest that reduced Nt-acetylation of one or more target substrates contributes to the pathogenesis of the Ogden syndrome.  相似文献   

19.
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   

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