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.
The N-glycosylation of the model nematode Caenorhabditis elegans has proven to be highly variable and rather complex; it is an example to contradict the existing impression that “simple” organisms possess also a rather simple glycomic capacity. In previous studies in a number of laboratories, N-glycans with up to four fucose residues have been detected. However, although the linkage of three fucose residues to the N,N′-diacetylchitobiosyl core has been proven by structural and enzymatic analyses, the nature of the fourth fucose has remained uncertain. By constructing a triple mutant with deletions in the three genes responsible for core fucosylation (fut-1, fut-6 and fut-8), we have produced a nematode strain lacking products of these enzymes, but still retaining maximally one fucose residue on its N-glycans. Using mass spectrometry and HPLC in conjunction with chemical and enzymatic treatments as well as NMR, we examined a set of α-mannosidase-resistant N-glycans. Within this glycomic subpool, we can reveal that the core β-mannose can be trisubstituted and so carries not only the ubiquitous α1,3- and α1,6-mannose residues, but also a “bisecting” β-galactose, which is substoichiometrically modified with fucose or methylfucose. In addition, the α1,3-mannose can also be α-galactosylated. Our data, showing the presence of novel N-glycan modifications, will enable more targeted studies to understand the biological functions and interactions of nematode glycans.Nematodes represent, along with arthropods, one of the largest groups of animals to exist on the planet; 25.000 species are described, but the existence of up to one million has been estimated (1, 2). They have various ecological niches and include free-living “worms” in the soil, fungivorous, entomopathogenic, and necromenic species as well as parasites of plants and mammals, which share the basic conserved body plan (more-or-less a digestive tube surrounded with muscle, whether larger or smaller). There are five major clades (Rhabditina, Enoplia, Spirurina, Tylenchina, and Dorylaimia) (2), yet the glycosylation of only a few nematode species has been studied with an inevitable focus on the model nematode Caenorhabditis elegans and parasitic species (3). Thereby, the use of C. elegans mutants has been highly valuable in dissecting aspects of nematode N-glycan biosynthesis and revealing the in vivo substrates for certain glycosyltransferases (4).As many nematodes are parasites, their interactions with the immune systems of their hosts have attracted attention; particularly, there are relationships between autoimmunity, allergy, vaccination, and helminth infections. The “old friends” hypothesis seeks to understand the evolutionary factors that have shaped the immune system and to explain correlations between lifestyles in the developed world and modern “epidemics,” which are due to immunological misbalance (57). Promising data have suggested that “worm therapy” may bring advantages to some patients with Crohn''s disease or allergies (8, 9); however, such approaches are controversial. Nevertheless, crude extracts even of Caenorhabditis elegans were shown to induce a glycan-dependent Th2 response (10), whereas the excretory-secretory products of some nematodes also have immunomodulatory activity (11). Furthermore, the native glycoproteins of some nematodes have proven effective in vaccination trials, whereas recombinant forms are not, which is suggestive that post-translational modifications may have a role in an efficacious immune response (12).As at least some of the molecules relevant to nematode immunomodulation or vaccination are glycoproteins, a proper understanding of nematode glycosylation is of biomedical and veterinary relevance. Over the years, it has become apparent that the core chitobiosyl region of nematode N-glycans is subject to a range of modifications, with up to three core fucose residues being present (α1,3- and α1,6-linked on the reducing-terminal “proximal” GlcNAc and α1,3-linked on the second “distal” GlcNAc). However, up to four fucose residues have been detected on C. elegans N-glycans and the exact nature of the linkage of the fourth fucose has remained obscure despite work in our own and other laboratories (3, 1315). Combined with the latest knowledge regarding the specificity of C. elegans core fucosyltransferases (13, 16, 17) as well as our recent data regarding the exact structures of N-glycans from the C. elegans double hexosaminidase mutant and other nematodes (1820), we concluded that some models for the tri- and tetrafucosylated N-glycans were incorrect. By preparing a triple mutant unable to core fucosylate its N-glycans, we generated a C. elegans strain containing maximally one fucose residue on the N-linked oligosaccharides. Thereby a pool of unusual mannosidase-resistant N-glycans was identified and, using mass spectrometry (MS) and NMR, we reveal their modification with bisecting galactose frequently capped with fucose or methylfucose.  相似文献   

2.
Phosphorylation is a widespread post-translational modification that modulates the function of a large number of proteins. Here we show that a significant proportion of all the domains in the human proteome is significantly enriched or depleted in phosphorylation events. A substantial improvement in phosphosites prediction is achieved by leveraging this observation, which has not been tapped by existing methods. Phosphorylation sites are often not shared between multiple occurrences of the same domain in the proteome, even when the phosphoacceptor residue is conserved. This is partly because of different functional constraints acting on the same domain in different protein contexts. Moreover, by augmenting domain alignments with structural information, we were able to provide direct evidence that phosphosites in protein-protein interfaces need not be positionally conserved, likely because they can modulate interactions simply by sitting in the same general surface area.Phosphorylation, the most widespread protein post-translational modification, is an important regulator of protein function. The addition of phosphate groups on serine, threonine, and tyrosine residues can modulate the activity of the target protein by inducing complex conformational changes, by modifying protein electrostatics, and by regulating domain-peptide interactions, as in 14-3-3 or SH2 domains, that specifically recognize phosphorylated residues. The standard experimental technique for the high-throughput identification of phosphorylation sites is mass spectrometry (1).Phosphorylation is catalyzed by protein kinases, a family that in humans comprises ∼540 members (2, 3). It is well understood that these enzymes recognize specific sequence motifs in their substrates (4, 5). Accordingly the sequence around the phosphorylation site is undisputedly the most important feature for phosphosite prediction (6, 7). However the “context,” in a broad sense, where these motifs occur is also important as sequence alone is not enough to achieve the observed specificity of phosphorylation. Therefore, several studies have characterized multiple aspects of phosphosites such as their preference for loops and disordered regions (reviewed in (8)), or the tendency of phosphoserines and phosphothreonines to occur in clusters (9), and these features have been used to improve the performance of phosphosite predictors (6, 7, 1012). Moreover placing kinases and substrates in the context of protein interaction networks has been shown to improve the prediction of phosphorylation by specific kinases (13).Perhaps one of the most puzzling observations when looking at the phosphoproteome as a whole, is the fact that a large proportion of phosphorylation sites is poorly conserved. This has led to various hypotheses. First some sites may represent nonfunctional, possibly low-stoichiometry, phosphorylation events that are picked up because of the sensitivity of mass-spectrometry (14, 15). Indeed functionally characterized sites and those matching known kinase motifs are more conserved on average (1517). However, although in biology function often equates with conservation, there could be genuinely functional fast-evolving phosphosites, that are responsible for species-specific differences in signaling and regulation. Moreover in some cases, especially in the regulation of protein-protein interactions, the exact position of the phosphosites may be unimportant (18, 19).Here we explore the issues of “context” and “conservation” of phosphorylation sites from the perspective of protein domains. To this end, we assembled a comprehensive database of phosphosites from publicly available sources and studied their proteome distribution with respect to the location and identity of protein domains. We focus on the human phosphoproteome because it has been very well characterized in a multitude of low- and high-throughput experiments, thus providing the opportunity for a comprehensive, proteome-wide, study. In particular, the issues we want to address are the following:
  1. Are specific domain types preferentially phosphorylated? Or conversely are some domains specifically depleted of phosphorylation sites?
  2. Can the domain context be used to improve the prediction of phosphorylation sites?
  3. What is the conservation pattern of phosphosites when looking at multiple instances of the same domain in the proteome?
  相似文献   

3.
4.
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   

5.
6.
The occurrence of highly conserved amyloid-forming sequences in Candida albicans Als proteins (H. N. Otoo et al., Eukaryot. Cell 7:776–782, 2008) led us to search for similar sequences in other adhesins from C. albicans and Saccharomyces cerevisiae. The β-aggregation predictor TANGO found highly β-aggregation-prone sequences in almost all yeast adhesins. These sequences had an unusual amino acid composition: 77% of their residues were β-branched aliphatic amino acids Ile, Thr, and Val, which is more than 4-fold greater than their prevalence in the S. cerevisiae proteome. High β-aggregation potential peptides from S. cerevisiae Flo1p and C. albicans Eap1p rapidly formed insoluble amyloids, as determined by Congo red absorbance, thioflavin T fluorescence, and fiber morphology. As examples of the amyloid-forming ability of the native proteins, soluble glycosylphosphatidylinositol (GPI)-less fragments of C. albicans Als5p and S. cerevisiae Muc1p also formed amyloids within a few days under native conditions at nM concentrations. There was also evidence of amyloid formation in vivo: the surfaces of cells expressing wall-bound Als1p, Als5p, Muc1p, or Flo1p were birefringent and bound the fluorescent amyloid-reporting dye thioflavin T. Both of these properties increased upon aggregation of the cells. In addition, amyloid binding dyes strongly inhibited aggregation and flocculation. The results imply that amyloid formation is an intrinsic property of yeast cell adhesion proteins from many gene families and that amyloid formation is an important component of cellular aggregation mediated by these proteins.Protein amyloids are characteristic of pathological conditions, including neurodegenerative diseases (4, 11, 17, 38). These protein aggregates can also occur naturally in adhesive bacterial curli (3), melanosomes (14), condensed peptide hormone arrays (24), as regulatory prions in yeast (2, 5), and fungal hydrophobins, which are nonantigenic coats to some fungi (1, 33, 39). Nevertheless, such natural occurrences are relatively few, considering the negative free energy for amyloid formation (28).We have recently discovered that there are amyloid-forming sequences in the cell surface Als adhesins of Candida albicans. Cells that express these adhesins aggregate readily, and the aggregation has amyloid-like properties, including protein conformational shifting, surface birefringence, and ability to bind the amyloid-active dyes Congo red and amino-naphthalene sulfonic acid (ANS) (29). A five- to seven-residue sequence in Als1p, Als3p, and Als5p has extremely high potential for formation of β-aggregates, according to the protein state prediction program TANGO (13, 27, 31). Such β-aggregates include amyloids, which are ordered structures with paracrystalline regions of stacked parallel β-strands that are perpendicular to the long axis of micrometer-long fibrils. The strands are stabilized by interaction of identical sequences from many protein molecules (31, 32). Where TANGO analyses have shown that specific sequences have β-aggregate potentials greater than 20%, an insoluble β-aggregate state is likely to form. These β-aggregates nucleate formation of amyloids if the proteins can associate to form fibers (13, 27, 31). Sequences in the conserved 127-residue T region of Als1p, Als3p, and Als5p have β-aggregation potentials of >90% (27). An oligopeptide with this sequence, as well as 412- and 645-residue fragments of Als5p formed authentic amyloids, as determined by characteristic dye binding and fiber morphology. The amyloid-forming sequences were rich in the β-branched amino acids Thr, Val, and Ile. This amino acid composition is unusual among proteins in general, but is common in the Thr-rich mid-piece domains of yeast adhesins.Yeasts display many cell-wall-bound adhesins that mediate colonial and biofilm interactions as well as host-pathogen binding (9, 21, 41). Such adhesins have a common mosaic structure. In general, the adhesins have N-terminal globular binding domains (often immunoglobulin-like or lectin-like), Thr-rich mid-piece sequences including tandem repeats, and 300- to 800-residue heavily glycosylated Ser and Thr-rich “stalk” domains near the C-terminal domain that extend the active regions from the surface of the wall. The adhesins are covalently cross-linked to wall polysaccharides through modified glycosylphosphatidylinositol (GPI) anchors and/or glycosyl esters of glutamic acid (9, 18).Because the yeast adhesins share this common modular domain structure, we searched among known and putative yeast adhesins for sequences with high β-aggregation potential. We have found that many of these proteins share amyloid-forming sequences and amyloid-like behavior on activation.  相似文献   

7.
Based on conventional data-dependent acquisition strategy of shotgun proteomics, we present a new workflow DeMix, which significantly increases the efficiency of peptide identification for in-depth shotgun analysis of complex proteomes. Capitalizing on the high resolution and mass accuracy of Orbitrap-based tandem mass spectrometry, we developed a simple deconvolution method of “cloning” chimeric tandem spectra for cofragmented peptides. Additional to a database search, a simple rescoring scheme utilizes mass accuracy and converts the unwanted cofragmenting events into a surprising advantage of multiplexing. With the combination of cloning and rescoring, we obtained on average nine peptide-spectrum matches per second on a Q-Exactive workbench, whereas the actual MS/MS acquisition rate was close to seven spectra per second. This efficiency boost to 1.24 identified peptides per MS/MS spectrum enabled analysis of over 5000 human proteins in single-dimensional LC-MS/MS shotgun experiments with an only two-hour gradient. These findings suggest a change in the dominant “one MS/MS spectrum - one peptide” paradigm for data acquisition and analysis in shotgun data-dependent proteomics. DeMix also demonstrated higher robustness than conventional approaches in terms of lower variation among the results of consecutive LC-MS/MS runs.Shotgun proteomics analysis based on a combination of high performance liquid chromatography and tandem mass spectrometry (MS/MS) (1) has achieved remarkable speed and efficiency (27). In a single four-hour long high performance liquid chromatography-MS/MS run, over 40,000 peptides and 5000 proteins can be identified using a high-resolution Orbitrap mass spectrometer with data-dependent acquisition (DDA)1 (2, 3). However, in a typical LC-MS analysis of unfractionated human cell lysate, over 100,000 individual peptide isotopic patterns can be detected (4), which corresponds to simultaneous elution of hundreds of peptides. With this complexity, a mass spectrometer needs to achieve ≥25 Hz MS/MS acquisition rate to fully sample all the detectable peptides, and ≥17 Hz to cover reasonably abundant ones (4). Although this acquisition rate is reachable by modern time-of-flight (TOF) instruments, the reported DDA identification results do not encompass all expected peptides. Recently, the next-generation Orbitrap instrument, working at 20 Hz MS/MS acquisition rate, demonstrated nearly full profiling of yeast proteome using an 80 min gradient, which opened the way for comprehensive analysis of human proteome in a time efficient manner (5).During the high performance liquid chromatography-MS/MS DDA analysis of complex samples, high density of co-eluting peptides results in a high probability for two or more peptides to overlap within an MS/MS isolation window. With the commonly used ±1.0–2.0 Th isolation windows, most MS/MS spectra are chimeric (4, 810), with cofragmenting precursors being naturally multiplexed. However, as has been discussed previously (9, 10), the cofragmentation events are currently ignored in most of the conventional analysis workflows. According to the prevailing assumption of “one MS/MS spectrum–one peptide,” chimeric MS/MS spectra are generally unwelcome in DDA, because the product ions from different precursors may interfere with the assignment of MS/MS fragment identities, increasing the rate of false discoveries in database search (8, 9). In some studies, the precursor isolation width was set as narrow as ±0.35 Th to prevent unwanted ions from being coselected, fragmented or detected (4, 5).On the contrary, multiplexing by cofragmentation is considered to be one of the solid advantages in data-independent acquisition (DIA) (1013). In several commonly used DIA methods, the precursor ion selection windows are set much wider than in DDA: from 25 Th as in SWATH (12), to extremely broad range as in AIF (13). In order to use the benefit of MS/MS multiplexing in DDA, several approaches have been proposed to deconvolute chimeric MS/MS spectra. In “alternative peptide identification” method implemented in Percolator (14), a machine learning algorithm reranks and rescores peptide-spectrum matches (PSMs) obtained from one or more MS/MS search engines. But the deconvolution in Percolator is limited to cofragmented peptides with masses differing from the target peptide by the tolerance of the database search, which can be as narrow as a few ppm. The “active demultiplexing” method proposed by Ledvina et al. (15) actively separates MS/MS data from several precursors using masses of complementary fragments. However, higher-energy collisional dissociation often produces MS/MS spectra with too few complementary pairs for reliable peptide identification. The “MixDB” method introduces a sophisticated new search engine, also with a machine learning algorithm (9). And the “second peptide identification” method implemented in Andromeda/MaxQuant workflow (16) submits the same dataset to the search engine several times based on the list of chromatographic peptide features, subtracting assigned MS/MS peaks after each identification round. This approach is similar to the ProbIDTree search engine that also performed iterative identification while removing assigned peaks after each round of identification (17).One important factor for spectral deconvolution that has not been fully utilized in most conventional workflows is the excellent mass accuracy achievable with modern high-resolution mass spectrometry (18). An Orbitrap Fourier-transform mass spectrometer can provide mass accuracy in the range of hundreds of ppb (parts per billion) for mass peaks with high signal-to-noise (S/N) ratio (19). However, the mass error of peaks with lower S/N ratios can be significantly higher and exceed 1 ppm. Despite this dependence of the mass accuracy from the S/N level, most MS and MS/MS search engines only allow users to set hard cut-off values for the mass error tolerances. Moreover, some search engines do not provide the option of choosing a relative error tolerance for MS/MS fragments. Such negligent treatment of mass accuracy reduces the analytical power of high accuracy experiments (18).Identification results coming from different MS/MS search engines are sometimes not consistent because of different statistical assumptions used in scoring PSMs. Introduction of tools integrating the results of different search engines (14, 20, 21) makes the data interpretation even more complex and opaque for the user. The opposite trend—simplification of MS/MS data interpretation—is therefore a welcome development. For example, an extremely straightforward algorithm recently proposed by Wenger et al. (22) demonstrated a surprisingly high performance in peptide identification, even though it is only marginally more complex than simply counting the number of matches of theoretical fragment peaks in high resolution MS/MS, without any a priori statistical assumption.In order to take advantage of natural multiplexing of MS/MS spectra in DDA, as well as properly utilize high accuracy of Orbitrap-based mass spectrometry, we developed a simple and robust data analysis workflow DeMix. It is presented in Fig. 1 as an expansion of the conventional workflow. Principles of some of the processes used by the workflow are borrowed from other approaches, including the custom-made mass peak centroiding (20), chromatographic feature detection (19, 20), and two-pass database search with the first limited pass to provide a “software lock mass” for mass scale recalibration (23).Open in a separate windowFig. 1.An overview of the DeMix workflow that expands the conventional workflow, shown by the dashed line. Processes are colored in purple for TOPP, red for search engine (Morpheus/Mascot/MS-GF+), and blue for in-house programs.In DeMix workflow, the deconvolution of chimeric MS/MS spectra consists of simply “cloning” an MS/MS spectrum if a potential cofragmented peptide is detected. The list of candidate peptide precursors is generated from chromatographic feature detection, as in the MaxQuant/Andromeda workflow (16, 19), but using The OpenMS Proteomics Pipeline (TOPP) (20, 24). During the cloning, the precursor is replaced by the new candidate, but no changes in the MS/MS fragment list are made, and therefore the cloned MS/MS spectra remain chimeric. Processing such spectra requires a search engine tolerant to the presence of unassigned peaks, as such peaks are always expected when multiple precursors cofragment. Thus, we chose Morpheus (22) as a search engine. Based on the original search algorithm, we implement a reformed scoring scheme: Morpheus-AS (advanced scoring). It inherits all the basic principles from Morpheus but deeper utilizes the high mass accuracy of the data. This kind of database search removes the necessity of spectral processing for physical separation of MS/MS data into multiple subspectra (15), or consecutive subtraction of peaks (16, 17).Despite the fact that DeMix workflow is largely a combination of known approaches, it provides remarkable improvement compared with the state-of-the-art. On our Orbitrap Q-Exactive workbench, testing on a benchmark dataset of two-hour single-dimension LC-MS/MS experiments from HeLa cell lysate, we identified on average 1.24 peptide per MS/MS spectrum, breaking the “one MS/MS spectrum–one peptide” paradigm on the level of whole data set. At 1% false discovery rate (FDR), we obtained on average nine PSMs per second (at the actual acquisition rate of ca. seven MS/MS spectra per second), and detected 40 human proteins per minute.  相似文献   

8.
9.
The Acanthamoeba castellanii myosin-Is were the first unconventional myosins to be discovered, and the myosin-I class has since been found to be one of the more diverse and abundant classes of the myosin superfamily. We used two-dimensional (2D) crystallization on phospholipid monolayers and negative stain electron microscopy to calculate a projection map of a “classical” myosin-I, Acanthamoeba myosin-IB (MIB), at ∼18 Å resolution. Interpretation of the projection map suggests that the MIB molecules sit upright on the membrane. We also used cryoelectron microscopy and helical image analysis to determine the three-dimensional structure of actin filaments decorated with unphosphorylated (inactive) MIB. The catalytic domain is similar to that of other myosins, whereas the large carboxy-terminal tail domain differs greatly from brush border myosin-I (BBM-I), another member of the myosin-I class. These differences may be relevant to the distinct cellular functions of these two types of myosin-I. The catalytic domain of MIB also attaches to F-actin at a significantly different angle, ∼10°, than BBM-I. Finally, there is evidence that the tails of adjacent MIB molecules interact in both the 2D crystal and in the decorated actin filaments.The myosin superfamily consists of at least 12 distinct classes that vary both in the sequence of their conserved myosin catalytic domains as well as in their unique tails (Mooseker and Cheney, 1995; Sellers and Goodson, 1995). For many years the only known myosins were the double-headed, filament-forming myosins found in muscle (conventional myosins or myosins-II). The remaining classes of myosin have been termed “unconventional myosins” to differentiate them from the myosins-II. Probably the most thoroughly studied class of unconventional myosins is the myosin-I class. These small, single-headed myosins bind to membrane lipids through a basic domain in their tail (for review see Pollard et al., 1991; Mooseker and Cheney, 1995). The first unconventional myosin (and first myosin-I) was isolated from Acanthamoeba castellanii (Pollard and Korn, 1973 a,b), and was purified on the basis of its K+, EDTA, and actin-activated MgATPase activities. However, this myosin was unusual in that it had a single heavy chain of ∼140 kD, in contrast to the two ∼200-kD heavy chains of myosin-II (Pollard and Korn, 1973 a).Three isoforms of the classical Acanthamoeba myosins-I are now known: myosins-IA, -IB, and -IC (Maruta and Korn, 1977a ,b; Maruta et al., 1979). Each of the isoforms consists of a conserved myosin catalytic domain, a binding site for one or two light chains, a basic domain, a GPA1(Q) domain (rich in glycine, proline and alanine [glutamine]), and an scr-homology domain-3 (SH3) domain (Pollard et al., 1991; Mooseker and Cheney, 1995). These myosins-I can associate with membranes or with actin filaments through their tail domains. An electrostatic association of myosin-I with anionic phospholipids and with base-stripped membranes has been shown to occur (Adams and Pollard, 1989; Miyata et al., 1989; Hayden et al., 1990), and this interaction has been mapped to the basic domain (Doberstein and Pollard, 1992). Interestingly, these myosins also contain a second, ATP-insensitive actin binding site (Lynch et al., 1986) enabling them to mediate actin–actin movements (Albanesi et al., 1985; Fujisaki et al., 1985). In myosin-IA (Lynch et al., 1986) and myosin-IC (Doberstein and Pollard, 1992), this binding site was localized to the GPA domain. Acanthamoeba myosins-I have maximal steady-state actin-activated ATPase rates of ∼10–20 s−1 (Pollard and Korn, 1973 b; Albanesi et al., 1983), and an unusual triphasic dependence upon actin concentration (Pollard and Korn, 1973 b; Albanesi et al., 1983). This triphasic activation is due to the actin cross-linking ability imparted by the ATP-insensitive actin binding site on the tail (Albanesi et al., 1985). Analysis of the individual steps in the ATPase cycle by transient kinetics revealed that the mechanism of myosin-IA is similar to slow skeletal muscle myosin, whereas myosin-IB (MIB) is similar to fast skeletal muscle myosin (Ostap and Pollard, 1996). The in vitro motility of myosin-I has also been well characterized (Zot et al., 1992). The maximal rate of filament sliding is ∼0.2 μm s−1. Interestingly, this rate is ∼10–50× slower than the rates observed for skeletal muscle myosin, even though the ATPase rates are comparable.MIB consists of a 125-kD heavy chain and a single 27-kD light chain (Maruta et al., 1979; Jung et al., 1987). This isoform is primarily associated with the plasma membrane as well as vacuolar membranes (Baines et al., 1992). MIB appears to be associated with the plasma membrane at sites of phagocytosis and was concentrated at the tips of pseudopodia (Baines et al., 1992). This localization suggests that MIB may be involved in membrane dynamics at the cell surface. MIB is regulated by heavy chain phosphorylation of serine 411 (Brzeska et al., 1989, 1990), which is located at the actin-binding site (Rayment et al., 1993). Similar to the myosin-I isoforms in Acanthamoeba, heavy chain phosphorylation results in >20-fold activation of the actin-activated myosin-I ATPase activity (Albanesi et al., 1983). This activation is not the result of changes in the binding of myosin-I to F-actin (Albanesi et al., 1983; Ostap and Pollard, 1996). The transient kinetic studies of Ostap and Pollard (1996) suggest that phosphorylation regulates the rate-limiting phosphate release step, the transition from weakly bound intermediates in rapid equilibrium with actin to strongly bound states, capable of sustaining force.Despite the extensive analysis of ameboid myosin-I biochemical properties and in vivo function, there is little structural information on these myosins. The only detailed structural information available for the myosins-I comes from recent electron microscopy studies on brush border myosin-I (BBM-I) (Jontes et al., 1995; Jontes and Milligan, 1997a ,b; Whittaker and Milligan, 1997), a structurally distinct myosin-I subtype. Therefore, we investigated the structure of a “classical,” ameboid-type myosin, Acanthamoeba MIB using electron microscopy. First, electron micrographs of negatively stained two-dimensional (2D) crystals were used to generate a projection map of MIB at ∼18 Å resolution. In addition, we used cryoelectron microscopy and helical image analysis to produce a moderate resolution three-dimensional (3D) map (30 Å) of actin filaments decorated with MIB. These studies enabled us to compare the structure of MIB with BBM-I. The comparison of MIB with BBM-I reveals marked structural differences in the tail domains of these two proteins; MIB appears to have a much shorter “lever arm” and a more compact tail, whereas most of the BBM-I mass is composed of an extended light chain–binding domain (LCBD). In addition, the MIB catalytic domain appears to be slightly tilted compared to BBM-I, with respect to the F-actin axis. Our structural results suggest that these two types of myosin-I may have distinct intracellular functions.  相似文献   

10.
In the last two years, because of advances in protein separation and mass spectrometry, top-down mass spectrometry moved from analyzing single proteins to analyzing complex samples and identifying hundreds and even thousands of proteins. However, computational tools for database search of top-down spectra against protein databases are still in their infancy. We describe MS-Align+, a fast algorithm for top-down protein identification based on spectral alignment that enables searches for unexpected post-translational modifications. We also propose a method for evaluating statistical significance of top-down protein identifications and further benchmark various software tools on two top-down data sets from Saccharomyces cerevisiae and Salmonella typhimurium. We demonstrate that MS-Align+ significantly increases the number of identified spectra as compared with MASCOT and OMSSA on both data sets. Although MS-Align+ and ProSightPC have similar performance on the Salmonella typhimurium data set, MS-Align+ outperforms ProSightPC on the (more complex) Saccharomyces cerevisiae data set.In the past two decades, proteomics was dominated by bottom-up mass spectrometry that analyzes digested peptides rather than intact proteins. Bottom-up approaches, although powerful, do have limitations in analyzing protein species, e.g. various proteolytic forms of the same protein or various protein isoforms resulting from alternative splicing. Top-down mass spectrometry focuses on analyzing intact proteins and large peptides (110) and has advantages in localizing multiple post-translational modifications (PTMs)1 in a coordinated fashion (e.g. combinatorial PTM code) and identifying multiple protein species (e.g. proteolytically processed protein species) (11). Until recently, most top-down studies were limited to single purified proteins (1215). Top-down studies of protein mixtures were restricted by difficulties in separating and fragmenting intact proteins and a shortage of robust computational tools.In the last two years, because of advances in protein separation and top-down instrumentation, top-down mass spectrometry moved from analyzing single proteins to analyzing complex samples containing hundreds and even thousands of proteins (1621). Because algorithms for interpreting top-down spectra are still in their infancy, many recent developments include computational innovations in protein identification.Because top-down spectra are complex, the first step in top-down spectral interpretation is usually spectral deconvolution, which converts a complex top-down spectrum to a list of monoisotopic masses (a deconvolved spectrum). Every protein (possibly with modifications) can be scored against a top-down deconvoluted spectrum, resulting in a Protein-Spectrum-Match (PrSM). The top-down protein identification problem is finding a protein in a database with the highest scoring PrSM for a top-down spectrum and further output the PrSM if it is statistically significant. There are several software tools for top-down protein identification (
SoftwareIdentification of unexpected modificationsProteogenomics search against 6-frame translationSpeedEstimation of statistical significance
ProSightPC+/−a+Fast/Slowb+
PIITA+/−Fast
UStag++Fast
MS-TopDown+Slow
MS-Align+++Fast+
Open in a separate windowa ProSightPC has various search modes that contribute to bridging the gap between blind and restrictive modes of MS/MS database search. It can identify truncated proteins by using biomarker search and identify unexpected modifications by using Δm mode and setting the error tolerance of precursor mass to a large value (e.g., 1999 Da). However, it is not designed for identifying truncated proteins with unexpected PTMs which are not represented in the “shotgun annotated” database.b In its most advances mode, ProSightPC can search the annotated top-down database that contains various protein species. However, ProSightPC searches in this mode become an order of magnitude slower.We describe MS-Align+, a fast software tool for top-down protein identification. MS-Align+ shares the spectral alignment approach with MS-TopDown, but greatly improves on speed, statistical analysis (providing E-values of PrSMs), and the number of identified PrSMs (e.g. by finding spectral alignments between spectra and truncated proteins). We benchmarked various tools for top-down protein identification on two data sets from Saccharomyces cerevisiae (SC) and Salmonella typhimurium (ST). We demonstrate that MS-Align+ significantly increase the number of identified spectra as compared with MASCOT and OMSSA on both data sets. Although MS-Align+ and ProSightPC have similar performance on the ST data set, MS-Align+ outperforms ProSightPC on the more complex SC data set.  相似文献   

11.
Structure and Function of Glycosylated Tandem Repeats from Candida albicans Als Adhesins     
Aaron T. Frank  Caleen B. Ramsook  Henry N. Otoo  Cho Tan  Gregory Soybelman  Jason M. Rauceo  Nand K. Gaur  Stephen A. Klotz  Peter N. Lipke 《Eukaryotic cell》2010,9(3):405-414
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   

12.
Targeting Lipid Esterases in Mycobacteria Grown Under Different Physiological Conditions Using Activity-based Profiling with Tetrahydrolipstatin (THL)     
Madhu Sudhan Ravindran  Srinivasa P. S. Rao  Xiamin Cheng  Ankit Shukla  Amaury Cazenave-Gassiot  Shao Q. Yao  Markus R. Wenk 《Molecular & cellular proteomics : MCP》2014,13(2):435-448
Tetrahydrolipstatin (THL) is bactericidal but its precise target spectrum is poorly characterized. Here, we used a THL analog and activity-based protein profiling to identify target proteins after enrichment from whole cell lysates of Mycobacterium bovis Bacillus Calmette-Guérin cultured under replicating and non-replicating conditions. THL targets α/β-hydrolases, including many lipid esterases (LipD, G, H, I, M, N, O, V, W, and TesA). Target protein concentrations and total esterase activity correlated inversely with cellular triacylglycerol upon entry into and exit from non-replicating conditions. Cellular overexpression of lipH and tesA led to decreased THL susceptibility thus providing functional validation. Our results define the target spectrum of THL in a biological species with particularly diverse lipid metabolic pathways. We furthermore derive a conceptual approach that demonstrates the use of such THL probes for the characterization of substrate recognition by lipases and related enzymes.Mycobacterium tuberculosis, the causative agent of tuberculosis (TB), is responsible for nearly 2 million deaths each year. The host immune response toward aerosol infection is to quarantine tubercle bacilli in a granulomatous structure (1, 2). However, granuloma-associated mycobacteria can switch to a non-replicative, “dormant” state and successfully evade immune response for decades after infection (3, 4). The metabolic events that permit tubercle bacilli to enter host cells and revive from states of persistence suggest that lipids are utilized as a carbon source (57). During times of oxygen deprivation and in the absence of host cells, cultivated mycobacteria store fatty acids (FAs) in the form of triacylglycerol (TAG)1-enriched lipid droplets (810). Upon resuscitation (by the re-introduction of oxygen), these lipid droplets vanish and TAGs are hydrolyzed (11). Unfortunately, the molecular mechanisms for TAG build-up and breakdown are far less well understood in bacteria when compared with those processes in eukaryotes.Comparative sequence analysis of the Mtb genome has revealed that it contains 250 genes encoding enzymes involved in lipid metabolism compared with only 50 enzymes in Escherichia coli, which has a genome of comparable size. Among these genes, 150 are predicted to encode proteins involved in lipid catabolism (12, 13). A family of 24 carboxyl ester hydrolases called “lip” genes (lipC to Z, except K and S) has been predicted to play a role in lipid catabolism (14). Among these, only a few have been functionally characterized and related to mycobacterial dormancy and resuscitation (1518).Tetrahydrolipstatin, a serine esterase inhibitor, covalently binds to and inhibits mammalian lipases and fatty acid synthase (FAS) and is marketed as “Orlistat” for the treatment of severe forms of obesity (19). THL was previously shown to inhibit both active and latent forms of mycobacteria (11, 2022) but the bacterial target spectrum remains poorly characterized. Therefore, to (1) define the THL target spectrum in a mycobacterial species and (2) to obtain biochemical insights into regulation of lipases and esterases in different metabolic states, we employed a chemical-proteomics approach using activity-based protein profiling (ABPP) with a bait that has been described to bind to lipolytic enzymes (2325). We identified several known lipases (as anticipated), putative lipase and esterases, and hypothetical proteins of unknown functions, thereby providing a comprehensive resource of experimentally determined THL targets in mycobacteria. Importantly, we systematically compared readouts of fluorescently tagged THL-proteins (7 bands on one-dimensional SDS-PAGE) with those of mass spectrometry-based peptide identification of enriched protein fractions (247 in growing cells). This comparison led to the identification of 14 THL targets, two of which were further validated experimentally. We furthermore provide a conceptual framework for the evaluation of this target list using both experimental as well as bioinformatics approaches in two examples, lipH and tesA. Overall, our data indicate that THL is an anti-mycobacterial drug because of its potential to (1) bind to a relatively wide range of lipolytic enzymes and (2) prevent bacilli from resuscitating from a nonreplicating persistent (NRP) state when lipid metabolism is particularly important.  相似文献   

13.
Initial Quantitative Proteomic Map of 28 Mouse Tissues Using the SILAC Mouse     
Tamar Geiger  Ana Velic  Boris Macek  Emma Lundberg  Caroline Kampf  Nagarjuna Nagaraj  Mathias Uhlen  Juergen Cox  Matthias Mann 《Molecular & cellular proteomics : MCP》2013,12(6):1709-1722
  相似文献   

14.
Quantitative Proteomics Reveals Factors Regulating RNA Biology as Dynamic Targets of Stress-induced SUMOylation in Arabidopsis     
Marcus J. Miller  Mark Scalf  Thérèse C. Rytz  Shane L. Hubler  Lloyd M. Smith  Richard D. Vierstra 《Molecular & cellular proteomics : MCP》2013,12(2):449-463
  相似文献   

15.
Transcriptional Regulation of Streptomyces coelicolor Pathway-Specific Antibiotic Regulators by the absA and absB Loci          下载免费PDF全文
David J. Aceti  Wendy C. Champness 《Journal of bacteriology》1998,180(12):3100-3106
  相似文献   

16.
Naegleria gruberi De Novo Basal Body Assembly Occurs via Stepwise Incorporation of Conserved Proteins     
Lillian K. Fritz-Laylin  Zoe June Assaf  Sean Chen  W. Zacheus Cande 《Eukaryotic cell》2010,9(6):860-865
Centrioles and basal bodies are discrete structures composed of a cylinder of nine microtubule triplets and associated proteins. Metazoan centrioles can be found at mitotic spindle poles and are called basal bodies when used to organize microtubules to form the core structure of flagella. Naegleria gruberi, a unicellular eukaryote, grows as an amoeba that lacks a cytoplasmic microtubule cytoskeleton. When stressed, Naegleria rapidly (and synchronously) differentiates into a flagellate, forming a complete cytoplasmic cytoskeleton de novo, including two basal bodies and flagella. Here, we show that Naegleria has genes encoding conserved centriole proteins. Using novel antibodies, we describe the localization of three centrosomal protein homologs (SAS-6, γ-tubulin, and centrin-1) during the assembly of the flagellate microtubule cytoskeleton. We also used these antibodies to show that Naegleria expresses the proteins in the same order as their incorporation into basal bodies, with SAS-6 localizing first, followed by centrin and finally γ-tubulin. The similarities between basal body assembly in Naegleria and centriole assembly in animals indicate that mechanisms of assembly, as well as structure, have been conserved throughout eukaryotic evolution.The beautiful and enigmatic pinwheel structures of centrioles and basal bodies have captured the imaginations of cell biologists for over a century. These small (∼1-μm) organelles are composed largely of a cylinder of nine microtubule triplets (11). The surrounding amorphous material harbors the microtubule-organizing activities of the centrosome, placing centrioles at the hub of the microtubule cytoskeleton. Metazoan centrosomes define mitotic spindle poles, and their centrioles are called basal bodies when used to form cilia (29). Moreover, in 1900 Meeves showed in a series of classical experiments that centrioles and basal bodies are interconvertible structures (34). Centrioles must replicate exactly once per cell cycle, as duplication errors can lead to problems with chromosome segregation and cell morphology (17).Virtually all animal cells have a pair of centrosomal centrioles that duplicate via “templated” assembly, with the new centriole developing perpendicular and attached to a preexisting centriole (4). Centrioles can also be formed “de novo” in cytosol devoid of preexisting centrioles and basal bodies (20). In addition to many in vivo examples (20), terminally differentiated fibroblasts held in S phase can assemble centrioles de novo after removal of preexisting centrioles by laser microsurgery (15).The amoeboflagellate Naegleria gruberi grows as an amoeba that completely lacks a cytoplasmic microtubule cytoskeleton. However, when exposed to stressors such as temperature, osmotic, or pH changes, Naegleria rapidly differentiates into a flagellate, forming a complete cytoplasmic cytoskeleton from scratch, including two basal bodies and flagella (8). This differentiation occurs synchronously, with approximately 90% of cells growing visible flagella in a 15-min window (T50 = 65 min after initiation of differentiation). As part of this differentiation, Naegleria has been shown to assemble the pinwheel structure of the basal bodies de novo, about 10 min before flagella are seen (11).Two centrosomal proteins that have been studied during Naegleria differentiation are centrin and γ-tubulin. Centrin is a calcium-binding phosphoprotein that is an integral component of the wall and lumen of basal bodies and of the pericentriolar lattice in many organisms (4, 19). During differentiation, Naegleria induces synthesis of centrin protein, which then localizes specifically to basal body structures throughout differentiation (18). γ-Tubulin is a general microtubule nucleation factor that localizes to microtubule-organizing centers (MTOCs) of many types. Surprisingly, Naegleria''s γ-tubulin homolog has been reported to localize to basal body precursor complexes and then move to the other end of the cell before disappearing completely (32).A third protein that has come under recent scrutiny for its role in centriole duplication is SAS-6, a functionally conserved coiled-coil protein required for the formation of diverse basal body precursor structures (7, 21,23, 31). In Caenorhabditis elegans and Drosophila melanogaster, SAS-6 is recruited at S phase to form the “central tube,” a cylindrical basal body precursor that lacks microtubules (22, 23). SAS-6 is also required for the formation of the flat ring or cartwheel with nine radiating spokes, which is the first structure to be formed in the Chlamydomonas basal body (21).To determine if Naegleria is likely to have typical basal body components, we identified conserved basal body genes in the Naegleria genome. We also made antibodies to and localized Naegleria''s homologs of SAS-6 and γ-tubulin. Finally, we have determined the order of expression and incorporation of these proteins, as well as centrin, during Naegleria de novo basal body assembly.  相似文献   

17.
Localization and Targeting of an Unusual Pyridine Nucleotide Transhydrogenase in Entamoeba histolytica     
Mohammad Abu Yousuf  Fumika Mi-ichi  Kumiko Nakada-Tsukui  Tomoyoshi Nozaki 《Eukaryotic cell》2010,9(6):926-933
Pyridine nucleotide transhydrogenase (PNT) catalyzes the direct transfer of a hydride-ion equivalent between NAD(H) and NADP(H) in bacteria and the mitochondria of eukaryotes. PNT was previously postulated to be localized to the highly divergent mitochondrion-related organelle, the mitosome, in the anaerobic/microaerophilic protozoan parasite Entamoeba histolytica based on the potential mitochondrion-targeting signal. However, our previous proteomic study of isolated phagosomes suggested that PNT is localized to organelles other than mitosomes. An immunofluorescence assay using anti-E. histolytica PNT (EhPNT) antibody raised against the NADH-binding domain showed a distribution to the membrane of numerous vesicles/vacuoles, including lysosomes and phagosomes. The domain(s) required for the trafficking of PNT to vesicles/vacuoles was examined by using amoeba transformants expressing a series of carboxyl-terminally truncated PNTs fused with green fluorescent protein or a hemagglutinin tag. All truncated PNTs failed to reach vesicles/vacuoles and were retained in the endoplasmic reticulum. These data indicate that the putative targeting signal is not sufficient for the trafficking of PNT to the vesicular/vacuolar compartments and that full-length PNT is necessary for correct transport. PNT displayed a smear of >120 kDa on SDS-PAGE gels. PNGase F and tunicamycin treatment, chemical degradation of carbohydrates, and heat treatment of PNT suggested that the apparent aberrant mobility of PNT is likely attributable to its hydrophobic nature. PNT that is compartmentalized to the acidic compartments is unprecedented in eukaryotes and may possess a unique physiological role in E. histolytica.Pyridine nucleotide transhydrogenase (PNT) participates in the bioenergetic processes of the cell. PNT generally resides on the cytoplasmic membranes of bacteria and the inner membrane of mammalian mitochondria (3, 16) and utilizes the electrochemical proton gradient across the membrane to drive NADPH formation from NADH (14, 15, 39) according to the reaction H+out + NADH + NADP+↔H+in + NAD+ + NADPH, where “out” and “in” denote the cytosol and the matrix of the mitochondria, or the periplasmic space and the cytosol of bacteria, respectively.PNT has been identified in several protozoan parasites, including Entamoeba histolytica (8, 51), Eimeria tenella (17, 47), Mastigamoeba balamuthi (11) Plasmodium falciparum (10), Plasmodium yoelii (6), and Plasmodium berghei (12). In general, PNT contains conserved structural units consisting of three domains, the NAD(H)-binding domain (domain I [dI]) and the NADP(H)-binding domain (domain III [dIII]), both of which face the matrix side of the eukaryotic mitochondria or the cytoplasmic side in bacteria, and the hydrophobic domain (domain II [dII]), containing 11 to 13 transmembrane regions. PNT from E. tenella and E. histolytica exists as a single polypeptide in an unusual configuration consisting of dIIb-dIII-dI-dIIa, with a 38-amino-acid-long linker region between dIII and dI (48).E. histolytica, previously considered an “amitochondriate” protist, is currently considered to possess a mitochondrion-related organelle with reduced and divergent functions, the mitosome (1, 21, 23a, 26, 42). Our recent proteomic study of isolated mitosomes identified about 20 new constituents (26), together with four proteins previously demonstrated in E. histolytica mitosomes: Cpn60 (8, 19, 21, 42), Cpn10 (46), mitochondrial Hsp70 (2, 44), and mitochondrion carrier family (MCF) (ADP/ATP transporter) (7). Despite the early presumption of PNT being localized in mitosomes (8), based on the amino-terminal region rich in hydroxylated (five serines and threonines) and acidic (three glutamates) amino acids, which slightly resembles known mitochondrion- and hydrogenosome-targeting sequences (8, 35), PNT was not discovered in the mitosome proteome. We also doubted this premise because PNT was one of the major proteins identified in isolated phagosomes (32, 33). Thus, the intracellular localization and trafficking of PNT remain unknown.In this report, we showed that E. histolytica PNT (EhPNT) is localized to various vesicles and vacuoles, including lysosomes and phagosomes, using wild-type amoebae and antiserum raised against recombinant EhPNT and an E. histolytica line expressing EhPNT with a carboxyl-terminal hemagglutinin (HA) epitope tag and anti-HA antibody. We also showed that all domains of EhPNT are required for its trafficking to the acidic compartment by using amoeba transformants expressing the HA tag or green fluorescent protein (GFP) fused with a region containing various domains of EhPNT.  相似文献   

18.
A Possible Role for Exocytosis in Aflatoxin Export in Aspergillus parasiticus     
Anindya Chanda  Ludmila V. Roze  John E. Linz 《Eukaryotic cell》2010,9(11):1724-1727
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   

19.
A Novel Point Mutation at Position 156 of Reverse Transcriptase from Feline Immunodeficiency Virus Confers Resistance to the Combination of (−)-β-2′,3′-Dideoxy-3′-Thiacytidine and 3′-Azido-3′-Deoxythymidine     
Robert A. Smith  Kathryn M. Remington  Bradley D. Preston  Raymond F. Schinazi  Thomas W. North 《Journal of virology》1998,72(3):2335
  相似文献   

20.
Preserved Proteins from Extinct Bison latifrons Identified by Tandem Mass Spectrometry; Hydroxylysine Glycosides are a Common Feature of Ancient Collagen     
Ryan C. Hill  Matthew J. Wither  Travis Nemkov  Alexander Barrett  Angelo D'Alessandro  Monika Dzieciatkowska  Kirk C. Hansen 《Molecular & cellular proteomics : MCP》2015,14(7):1946-1958
Bone samples from several vertebrates were collected from the Ziegler Reservoir fossil site, in Snowmass Village, Colorado, and processed for proteomics analysis. The specimens come from Pleistocene megafauna Bison latifrons, dating back ∼120,000 years. Proteomics analysis using a simplified sample preparation procedure and tandem mass spectrometry (MS/MS) was applied to obtain protein identifications. Several bioinformatics resources were used to obtain peptide identifications based on sequence homology to extant species with annotated genomes. With the exception of soil sample controls, all samples resulted in confident peptide identifications that mapped to type I collagen. In addition, we analyzed a specimen from the extinct B. latifrons that yielded peptide identifications mapping to over 33 bovine proteins. Our analysis resulted in extensive fibrillar collagen sequence coverage, including the identification of posttranslational modifications. Hydroxylysine glucosylgalactosylation, a modification thought to be involved in collagen fiber formation and bone mineralization, was identified for the first time in an ancient protein dataset. Meta-analysis of data from other studies indicates that this modification may be common in well-preserved prehistoric samples. Additional peptide sequences from extracellular matrix (ECM) and non-ECM proteins have also been identified for the first time in ancient tissue samples. These data provide a framework for analyzing ancient protein signatures in well-preserved fossil specimens, while also contributing novel insights into the molecular basis of organic matter preservation. As such, this analysis has unearthed common posttranslational modifications of collagen that may assist in its preservation over time. The data are available via ProteomeXchange with identifier PXD001827.During the last decade, paleontology and taphonomy (the study of decaying organisms over time and the fossilization processes) have begun to overlap with the field of proteomics to shed new light on preserved organic matter in fossilized bones (14). These bones represent a time capsule of ancient biomolecules, owing to their natural resistance to post mortem decay arising from a unique combination of mechanical, structural, and chemical properties (47).Although bones can be cursorily described as a composite of collagen (protein) and hydroxyapatite (mineral), fossilized bones undergo three distinct diagenesis pathways: (i) chemical deterioration of the organic phase; (ii) chemical deterioration of the mineral phase; and (iii) (micro)biological attack of the composite (6). In addition, the rate of these degradation pathways are affected by temperature, as higher burial temperatures have been shown to accelerate these processes (6, 8). Though relatively unusual, the first of these three pathways results in a slower deterioration process, which is more generally mitigated under (6) specific environmental constraints, such as geochemical stability (stable temperature and acidity) that promote bone mineral preservation. Importantly, slower deterioration results in more preserved biological materials that are more amenable to downstream analytical assays. One example of this is the controversial case of bone and soft-tissue preservation from the Cretaceous/Tertiary boundary (922). In light of these and other studies of ancient biomolecules, paleontological models have proposed that organic biomolecules in ancient samples, such as collagen sequences from the 80 million-year-(my)-old Campanian hadrosaur, Brachylophosaurus canadensis (16) or 68-my-old Tyrannosaurus rex, might be protected by the microenvironment within bones. Such spaces are believed to form a protective shelter that is able to reduce the effects of diagenetic events. In addition to collagen, preserved biomolecules include blood proteins, cellular lipids, and DNA (4, 5). While the maximum estimated lifespan of DNA in bones is ∼20,000 years (ky) at 10 °C, bone proteins have an even longer lifespan, making them an exceptional target for analysis to gain relevant insights into fossilized samples (6). Indeed, the survival of collagen, which is considered to be the most abundant bone protein, is estimated to be in the range 340 ky at 20 °C. Similarly, osteocalcin, the second-most abundant bone protein, can persist for ≈45 ky at 20 °C, thus opening an unprecedented analytical window to study extremely old samples (2, 4, 23).Although ancient DNA amplification and sequencing can yield interesting clues and potential artifacts from contaminating agents (7, 2428), the improved preservation of ancient proteins provides access to a reservoir of otherwise unavailable genetic information for phylogenetic inference (25, 29, 30). In particular, mass spectrometry (MS)-based screening of species-specific collagen peptides has recently been used as a low-cost, rapid alternative to DNA sequencing for taxonomic attribution of morphologically unidentifiable small bone fragments and teeth stemming from diverse archeological contexts (25, 3133).For over five decades, researchers have presented biochemical evidence for the existence of preserved protein material from ancient bone samples (3436). One of the first direct measurements was by amino acid analysis, which showed that the compositional profile of ancient samples was consistent with collagens in modern bone samples (3739). Preservation of organic biomolecules, either from bone, dentin, antlers, or ivory, has been investigated by radiolabeled 14C fossil dating (40) to provide an avenue of delineating evolutionary divergence from extant species (3, 41, 42). It is also important to note that these parameters primarily depend on ancient bone collagen as the levels remain largely unchanged (a high percentage of collagen is retained, as gleaned by laboratory experiments on bone taphonomy (6)). Additionally, antibody-based immunostaining methods have given indirect evidence of intact peptide amide bonds (4345) to aid some of the first evidence of protein other than collagen and osteocalcin in ancient mammoth (43) and human specimens (46).In the past, mass spectrometry has been used to obtain MS signals consistent with modern osteocalcin samples (2, 47), and eventually postsource decay peptide fragmentation was used to confirm the identification of osteocalcin in fossil hominids dating back ∼75 ky (48). More recently, modern “bottom-up” proteomic methods were applied to mastodon and T. rex samples (10), complementing immunohistochemistry evidence (13, 17). The results hinted at the potential of identifying peptides from proteolytic digest of well-preserved bone samples. This work also highlighted the importance of minimizing sources of protein contamination and adhering to data publication guidelines (20, 21). In the past few years, a very well-preserved juvenile mammoth referred to as Lyuba was discovered in the Siberian permafrost and analyzed using high-resolution tandem mass spectrometry (29). This study was followed with a report by Wadsworth and Buckley (30) describing the analysis of proteins from 19 bovine bone samples spanning 4 ky to 1.5 my. Both of these groups reported the identification of additional collagen and noncollagen proteins.Recently, a series of large extinct mammal bones were unearthed at a reservoir near Snowmass Village, Colorado, USA (49, 50). The finding was made during a construction project at the Ziegler Reservoir, a fossil site that was originally a lake formed at an elevation of ∼2,705 m during the Bull Lake glaciations ∼140 ky ago (49, 51). The original lake area was ∼5 hectares in size with a total catchment of ∼14 hectares and lacked a direct water flow inlet or outlet. This closed drainage basin established a relatively unique environment that resulted in the exceptional preservation of plant material, insects (52), and vertebrate bones (49). In particular, a cranial specimen from extinct Bison latifrons was unearthed from the Biostratigraphic Zone/Marine Oxygen Isotope Stage (MIS) 5d, which dates back to ∼120 ky (53, 54).Here, we describe the use of paleoproteomics, for the identification of protein remnants with a focus on a particularly unique B. latifrons cranial specimen found at the Ziegler site. We developed a simplified sample processing approach that allows for analysis of low milligram quantities of ancient samples for peptide identification. Our method avoids the extensive demineralization steps of traditional protocols and utilizes an acid labile detergent to allow for efficient extraction and digestion without the need for additional sample cleanup steps. This approach was applied to a specimen from B. latifrons that displayed visual and mechanical properties consistent with the meninges, a fibrous tissue that lines the cranial cavity. Bioinformatics analysis revealed the presence of a recurring glycosylation signature in well-preserved collagens. In particular, the presence of glycosylated hydroxylysine residues was identified as a unique feature of bone fossil collagen, as gleaned through meta-analyses of raw data from previous reports on woolly mammoth (Mammuthus primigenius) and bovine samples (29, 30). The results from these meta-analyses indicate a common, unique feature of collagen that coincides with, and possibly contributes to its preservation.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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