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.
3.
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   

4.
5.
6.
7.
8.
Posttranslational modifications of proteins increase the complexity of the cellular proteome and enable rapid regulation of protein functions in response to environmental changes. Protein ubiquitylation is a central regulatory posttranslational modification that controls numerous biological processes including proteasomal degradation of proteins, DNA damage repair and innate immune responses. Here we combine high-resolution mass spectrometry with single-step immunoenrichment of di-glycine modified peptides for mapping of endogenous putative ubiquitylation sites in murine tissues. We identify more than 20,000 unique ubiquitylation sites on proteins involved in diverse biological processes. Our data reveals that ubiquitylation regulates core signaling pathways common for each of the studied tissues. In addition, we discover that ubiquitylation regulates tissue-specific signaling networks. Many tissue-specific ubiquitylation sites were obtained from brain highlighting the complexity and unique physiology of this organ. We further demonstrate that different di-glycine-lysine-specific monoclonal antibodies exhibit sequence preferences, and that their complementary use increases the depth of ubiquitylation site analysis, thereby providing a more unbiased view of protein ubiquitylation.Ubiquitin is a small 76-amino-acid protein that is conjugated to the ε-amino group of lysines in a highly orchestrated enzymatic cascade involving ubiquitin activating (E1), ubiquitin conjugating (E2), and ubiquitin ligase (E3) enzymes (1). Ubiquitylation is involved in the regulation of diverse cellular processes including protein degradation (2, 3, 4), DNA damage repair (5, 6), DNA replication (7), cell surface receptor endocytosis, and innate immune signaling (8, 9). Deregulation of protein ubiquitylation is implicated in the development of cancer and neurodegenerative diseases (10, 11). Inhibitors targeting the ubiquitin proteasome system are used in the treatment of hematologic malignancies such as multiple myeloma (12, 13).Recent developments in the mass spectrometry (MS)-based proteomics have greatly expedited proteome-wide analysis of posttranslational modifications (PTMs) (1417). Large-scale mapping of ubiquitylation sites by mass spectrometry is based on the identification of the di-glycine remnant that results from trypsin digestion of ubiquitylated proteins and remains attached to ubiquitylated lysines (18). Recently, two monoclonal antibodies were developed that specifically recognize di-glycine remnant modified peptides enabling their efficient enrichment from complex peptide mixtures (19, 20). These antibodies have been used to identify thousands of endogenous ubiquitylation sites in human cells, and to quantify site-specific changes in ubiquitylation in response to different cellular perturbations (2022). It should be noted that the di-glycine remnant is not specific for proteins modified by ubiquitin but also proteins modified by NEDD8 or ISG15 generate an identical di-glycine remnant on modified lysines making it impossible to distinguish between these modifications by mass spectrometry. However, expression of NEDD8 in mouse tissues was shown to be developmentally down-regulated (23), and ISG15 expression in bovine tissues is low in the absence of interferon stimulation (24). In cell culture experiments it was shown that a great majority of sites identified using di-glycine-lysine-specific antibodies stems from ubiquitylated peptides (20).The rates of cell proliferation and protein turnover in mammals vary dramatically between different tissues. Immortalized cell lines, often derived from cancer, are selected for high proliferation rates and fail to represent the complex conditions in tissues. Tissue proteomics can help to gain a more comprehensive understanding of physiological processes in multicellular organisms. Analysis of tissue proteome and PTMs can provide important insights into tissue-specific processes and signaling networks that regulate these processes (2532). In addition, development of mass spectrometric methods for analysis of PTMs in diseased tissues might lead to the identification of biomarkers.In this study, we combined high-resolution mass spectrometry with immunoenrichment of di-glycine modified peptides to investigate endogenous ubiquitylation sites in murine tissues. We identified more than 20,000 ubiquitylation sites from five different murine tissues and report the largest ubiquitylation dataset obtained from mammalian tissues to date. Furthermore, we compared the performance of the two monoclonal di-glycine-lysine-specific antibodies available for enrichment of ubiquitylated peptides, and reveal their relative preferences for different amino acids flanking ubiquitylation sites.  相似文献   

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

10.
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   

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

14.
A complete understanding of the biological functions of large signaling peptides (>4 kDa) requires comprehensive characterization of their amino acid sequences and post-translational modifications, which presents significant analytical challenges. In the past decade, there has been great success with mass spectrometry-based de novo sequencing of small neuropeptides. However, these approaches are less applicable to larger neuropeptides because of the inefficient fragmentation of peptides larger than 4 kDa and their lower endogenous abundance. The conventional proteomics approach focuses on large-scale determination of protein identities via database searching, lacking the ability for in-depth elucidation of individual amino acid residues. Here, we present a multifaceted MS approach for identification and characterization of large crustacean hyperglycemic hormone (CHH)-family neuropeptides, a class of peptide hormones that play central roles in the regulation of many important physiological processes of crustaceans. Six crustacean CHH-family neuropeptides (8–9.5 kDa), including two novel peptides with extensive disulfide linkages and PTMs, were fully sequenced without reference to genomic databases. High-definition de novo sequencing was achieved by a combination of bottom-up, off-line top-down, and on-line top-down tandem MS methods. Statistical evaluation indicated that these methods provided complementary information for sequence interpretation and increased the local identification confidence of each amino acid. Further investigations by MALDI imaging MS mapped the spatial distribution and colocalization patterns of various CHH-family neuropeptides in the neuroendocrine organs, revealing that two CHH-subfamilies are involved in distinct signaling pathways.Neuropeptides and hormones comprise a diverse class of signaling molecules involved in numerous essential physiological processes, including analgesia, reward, food intake, learning and memory (1). Disorders of the neurosecretory and neuroendocrine systems influence many pathological processes. For example, obesity results from failure of energy homeostasis in association with endocrine alterations (2, 3). Previous work from our lab used crustaceans as model organisms found that multiple neuropeptides were implicated in control of food intake, including RFamides, tachykinin related peptides, RYamides, and pyrokinins (46).Crustacean hyperglycemic hormone (CHH)1 family neuropeptides play a central role in energy homeostasis of crustaceans (717). Hyperglycemic response of the CHHs was first reported after injection of crude eyestalk extract in crustaceans. Based on their preprohormone organization, the CHH family can be grouped into two sub-families: subfamily-I containing CHH, and subfamily-II containing molt-inhibiting hormone (MIH) and mandibular organ-inhibiting hormone (MOIH). The preprohormones of the subfamily-I have a CHH precursor related peptide (CPRP) that is cleaved off during processing; and preprohormones of the subfamily-II lack the CPRP (9). Uncovering their physiological functions will provide new insights into neuroendocrine regulation of energy homeostasis.Characterization of CHH-family neuropeptides is challenging. They are comprised of more than 70 amino acids and often contain multiple post-translational modifications (PTMs) and complex disulfide bridge connections (7). In addition, physiological concentrations of these peptide hormones are typically below picomolar level, and most crustacean species do not have available genome and proteome databases to assist MS-based sequencing.MS-based neuropeptidomics provides a powerful tool for rapid discovery and analysis of a large number of endogenous peptides from the brain and the central nervous system. Our group and others have greatly expanded the peptidomes of many model organisms (3, 1833). For example, we have discovered more than 200 neuropeptides with several neuropeptide families consisting of as many as 20–40 members in a simple crustacean model system (5, 6, 2531, 34). However, a majority of these neuropeptides are small peptides with 5–15 amino acid residues long, leaving a gap of identifying larger signaling peptides from organisms without sequenced genome. The observed lack of larger size peptide hormones can be attributed to the lack of effective de novo sequencing strategies for neuropeptides larger than 4 kDa, which are inherently more difficult to fragment using conventional techniques (3437). Although classical proteomics studies examine larger proteins, these tools are limited to identification based on database searching with one or more peptides matching without complete amino acid sequence coverage (36, 38).Large populations of neuropeptides from 4–10 kDa exist in the nervous systems of both vertebrates and invertebrates (9, 39, 40). Understanding their functional roles requires sufficient molecular knowledge and a unique analytical approach. Therefore, developing effective and reliable methods for de novo sequencing of large neuropeptides at the individual amino acid residue level is an urgent gap to fill in neurobiology. In this study, we present a multifaceted MS strategy aimed at high-definition de novo sequencing and comprehensive characterization of the CHH-family neuropeptides in crustacean central nervous system. The high-definition de novo sequencing was achieved by a combination of three methods: (1) enzymatic digestion and LC-tandem mass spectrometry (MS/MS) bottom-up analysis to generate detailed sequences of proteolytic peptides; (2) off-line LC fractionation and subsequent top-down MS/MS to obtain high-quality fragmentation maps of intact peptides; and (3) on-line LC coupled to top-down MS/MS to allow rapid sequence analysis of low abundance peptides. Combining the three methods overcomes the limitations of each, and thus offers complementary and high-confidence determination of amino acid residues. We report the complete sequence analysis of six CHH-family neuropeptides including the discovery of two novel peptides. With the accurate molecular information, MALDI imaging and ion mobility MS were conducted for the first time to explore their anatomical distribution and biochemical properties.  相似文献   

15.
16.
17.
Mitotic division requires highly regulated morphological and biochemical changes to the cell. Upon commitment to exit mitosis, cells begin to remove mitotic regulators in a temporally and spatially controlled manner to bring about the changes that reestablish interphase. Ubiquitin-dependent pathways target these regulators to generate polyubiquitin-tagged substrates for degradation by the 26S proteasome. However, the lack of cell-based assays to investigate in vivo ubiquitination limits our knowledge of the identity of substrates of ubiquitin-mediated regulation in mitosis. Here we report an in vivo ubiquitin tagging system used in human cells that allows efficient purification of ubiquitin conjugates from synchronized cell populations. Coupling purification with mass spectrometry, we have identified a series of mitotic regulators targeted for polyubiquitination in mitotic exit. We show that some are new substrates of the anaphase-promoting complex/cyclosome and validate KIFC1 and RacGAP1/Cyk4 as two such targets involved respectively in timely mitotic spindle disassembly and cell spreading. We conclude that in vivo biotin tagging of ubiquitin can provide valuable information about the role of ubiquitin-mediated regulation in processes required for rebuilding interphase cells.Ubiquitination has emerged as a major post-translational modification determining the fate of cellular proteins. One of these fates is proteolysis, whereby the assembly of polyubiquitin chains creates signatures on target proteins that specify delivery to the 26S proteasome for proteolytic destruction. Targeted proteolysis is critical to the control of cell division. For example, the universally conserved mechanism of mitotic exit depends upon rapid proteolysis of mitotic cyclins and securins to drive the transition from mitosis to interphase. This transition is under surveillance by the spindle assembly checkpoint (SAC),1 which controls the activity of a multi-subunit ubiquitin ligase, the anaphase-promoting complex/cyclosome (APC/C) (1, 2).Much of the known specificity in the ubiquitin-proteasome system (UPS) is mediated at the level of substrate targeting by ubiquitin ligase (E3) enzymes, of which there are more than 600 in human cells. Given these facts, it is perhaps surprising that the APC/C is almost the only known engineer of the protein landscape after anaphase onset, targeting mitotic regulators for destruction with high temporal specificity (24). Some roles for nondegradative ubiquitination in regulating the localization of mitotic kinases Aurora B and Plk1 have been described (59), and a growing list of reported ubiquitin interactors can modulate ubiquitin-dependent events during mitosis (10). However, the majority of ubiquitination events that have so far been described as occurring at the transition from mitosis to interphase are APC/C-dependent.Two co-activator subunits, Cdc20 and Cdh1, play vital roles in APC/C-dependent substrate recognition (11) by recognizing two widely characterized degrons, the D-box and the KEN motif (12, 13). Computational approaches that have been used to calculate the total number of APC/C substrates from the prevalence of degrons in the human proteome estimate that there are between 100 and 200 substrates (14), and experiments using in vitro ubiquitination of protein arrays have given rise to estimates in the same range (15). Most of the mitotic regulators targeted by the APC/C during mitotic exit in human cells have been identified via in vitro degradation assays or ubiquitination assays on in vitro–expressed pools of substrates (1518). These approaches have identified several important substrates, but in the absence of in vivo parameters they may not identify substrates whose targeting depends on post-translational modifications or substrates that are only recognized in vivo as components of higher-order complexes. Not all substrates identified in this way have been validated as polyubiquitinated proteins in vivo. Multiple recent proteomic studies have identified large numbers of in vivo ubiquitin-modified sites from yeast (1921) and human cells (2229). None of these studies have used synchronized cell populations to provide information on the timing or regulation of substrate ubiquitination.We reasoned that a better view of ubiquitin-mediated processes that regulate mitotic exit would come from identifying proteins that are ubiquitinated in vivo during mitotic exit. With this goal in mind we adopted a system for in vivo tagging of ubiquitin chains with biotin, previously used to identify ubiquitin-conjugated proteins from the Drosophila neural system (30), and applied it to a human cell line (U2OS) that can be tightly synchronized at mitosis. In contrast to several recent studies that employed antibodies specific to the diGly-Lys remnant that marks ubiquitination sites following trypsin digestion (19, 25), an in vivo ubiquitin tagging strategy allows direct validation of candidate ubiquitinated proteins (whether mono- or polyubiquitinated) through immunoblotting of samples. Moreover, in contrast to other methods for affinity tagging of ubiquitin, or affinity purification via ubiquitin-binding domains, the use of the biotin tag enables purification under highly denaturing conditions for stringent isolation of ubiquitin-conjugated material from higher eukaryotes. His6-tagged ubiquitin is also available for use under denaturing conditions, but it is not generally useful in higher eukaryotic cells, where a high frequency of proteins containing multiple histidine residues confounds the specificity of nickel-affinity pulldowns (as discussed in detail in Ref. 30). Therefore, in this paper we describe the reproducible identification and validation of mitoticphase-specific polyubiquitinated proteins via the in vivo biotinylation of ubiquitin. A large number of polyubiquitinated proteins that we identified are specific to mitotic exit, when the APC/C is active, and we expect that many of them are substrates for the APC/C. We formally identified KIFC1/HSET and Cyk4/RACGAP1 as targets of APC/C-dependent ubiquitin-mediated proteolysis after anaphase onset and investigated the role of their ubiquitination in the regulation of mitotic exit. Cell cycle phase-specific information on protein ubiquitination and the generation of ubiquitinated protein networks provides a framework for further investigation of ubiquitin-controlled processes occurring during the rebuilding of interphase cells.  相似文献   

18.
19.
20.
Drug resistance poses a major challenge to ovarian cancer treatment. Understanding mechanisms of drug resistance is important for finding new therapeutic targets. In the present work, a cisplatin-resistant ovarian cancer cell line A2780-DR was established with a resistance index of 6.64. The cellular accumulation of cisplatin was significantly reduced in A2780-DR cells as compared with A2780 cells consistent with the general character of drug resistance. Quantitative proteomic analysis identified 340 differentially expressed proteins between A2780 and A2780-DR cells, which involve in diverse cellular processes, including metabolic process, cellular component biogenesis, cellular processes, and stress responses. Expression levels of Ras-related proteins Rab 5C and Rab 11B in A2780-DR cells were lower than those in A2780 cells as confirmed by real-time quantitative PCR and Western blotting. The short hairpin (sh)RNA-mediated knockdown of Rab 5C in A2780 cells resulted in markedly increased resistance to cisplatin whereas overexpression of Rab 5C in A2780-DR cells increases sensitivity to cisplatin, demonstrating that Rab 5C-dependent endocytosis plays an important role in cisplatin resistance. Our results also showed that expressions of glycolytic enzymes pyruvate kinase, glucose-6-phosphate isomerase, fructose-bisphosphate aldolase, lactate dehydrogenase, and phosphoglycerate kinase 1 were down-regulated in drug resistant cells, indicating drug resistance in ovarian cancer is directly associated with a decrease in glycolysis. Furthermore, it was found that glutathione reductase were up-regulated in A2780-DR, whereas vimentin, HSP90, and Annexin A1 and A2 were down-regulated. Taken together, our results suggest that drug resistance in ovarian cancer cell line A2780 is caused by multifactorial traits, including the down-regulation of Rab 5C-dependent endocytosis of cisplatin, glycolytic enzymes, and vimentin, and up-regulation of antioxidant proteins, suggesting Rab 5C is a potential target for treatment of drug-resistant ovarian cancer. This constitutes a further step toward a comprehensive understanding of drug resistance in ovarian cancer.Ovarian cancer is the major cause of death in women with gynecological cancer. Early diagnosis of ovarian cancer is difficult, while its progression is fast. The standard treatment is surgical removal followed by platinum-taxane chemotherapy. However, the efficacy of the traditional surgery and chemotherapy is rather compromised and platinum resistant cancer recurs in ∼25% of patients within six months, and the overall five-year survival rate is about 31% (13). Virtually no efficient second line treatment is available. In order to increase survival rates from ovarian cancer and enhance patients'' quality of life, new therapeutic targets are urgently required, necessitating a deeper understanding of molecular mechanisms of drug resistance.Mechanisms of drug-resistance in ovarian cancer have been extensively studied over the last 30 years. Earlier studies have found that multiple factors are linked to drug resistance in human ovarian cancer including reduced intracellular drug accumulation, intracellular cisplatin inactivation, and increased DNA repair (4). Reduced cellular drug accumulation is mediated by the copper transporter-1 responsible for the influx of cisplatin (59) and MDR1, which encodes an integral membrane protein named P-glycoprotein for the active efflux of platinum drugs. Up-regulation of MDR1 has been observed in cisplatin-treated ovarian cancer cells although cisplatin is not a substrate of P-glycoprotein (1013). A fraction of intracellular cisplatin can be converted into cisplatin-thiol conjugates by glutathione-S-transferase (GST) π, leading to inactivation of cisplatin. Up-regulation of both GSTπ and γ-glutamylcysteine synthetase has been associated with cisplatin resistance in ovarian, cervical and lung cancer cell lines (1418). Binding of cisplatin to DNA leads to intrastrand or interstrand cross-links that alter the structure of the DNA molecule causing DNA damage. It has been amply documented that pathways for recognition and repair of damaged DNA are up-regulated in drug-resistant cancer cells (1926). Furthermore, the secondary mutations have been identified, which restore the wild-type BRCA2 reading frame enhancing the acquired resistance to platinum-based chemotherapy (24). Alternations in other signaling pathways have also been found in drug resistant ovarian cancer (2729). For example, DNA-PK phosphorylates RAC-alpha serine/threonine-protein kinase (AKT) and inhibits cisplatin-mediated apoptosis (28); and silencing of HDAC4 increases acetyl-STAT1 levels to prevent platinum-induced STAT1 activation and restore cisplatin sensitivity (29).Proteomics is playing an increasingly important role in identifying differentially expressed proteins between drug-resistant and drug sensitive ovarian cancer cells (3035). An earlier study has identified 57 differentially expressed proteins in human ovarian cancer cells and their platinum-resistant sublines, including annexin A3, destrin, cofilin 1, Glutathione-S-transferase omega 1, and cytosolic NADP+-dependent isocitrate dehydrogenase using 2D gel electrophoresis (30). Employing a similar 2D gel electrophoresis approach, changes in protein expressions of capsid glycoprotein, fructose-bisphosphate aldolase C, heterogeneous nuclear ribonucleoproteins A2/B1, putative RNA-binding protein 3, Ran-specific GTPase-activating protein, ubiquitin carboxyl-terminal hydrolase isozyme L1, stathmin, ATPSH protein, chromobox protein homolog3, and phosphoglycerate kinase 1 (PGK)1 were found in A2780 and drug-resistant A2780 cells (32). It is worth mentioning that ALDO and PGK are glycolytic enzymes, indicating that glycolysis plays a role in drug resistance. Studies have demonstrated that resistance to platinum drugs in ovarian cancer cells is linked to mitochondrial dysfunctions in oxidative phosphorylation and energy production (3640). Mitochondrial proteomic analysis of drug-resistant cells has shown that five mitochondrial proteins (ATP-a, PRDX3, PHB, ETF, and ALDH) that participate in the electron transport respiratory chain are down-regulated in drug-resistant cell lines (41). PRDX3 is involved in redox regulation of the cell to protect radical-sensitive enzymes from oxidative damage. However, it is not clear how down-regulation of PRDX3 is associated with drug-resistance. A more recent study showed that activated leukocyte cell adhesion molecule (ALCA) and A kinase anchoring protein 12 (AKAP12) are elevated in drug-resistant A2780-CP20 cells by quantifying the mitochondrial proteins (42). Despite these efforts, the drug-resistance mechanisms are not yet well understood.In this work, we established and characterized a drug-resistant cell line A2780-DR from A2780 cells. We employed a quantitative proteomic method to identify the differentially expressed proteins between A2780 and A2780-DR cells. Expression changes of selected proteins were confirmed by qPCR and Western blotting. We also used shRNA silencing to explore functions of Rab 5C and Rab 11B proteins in drug resistance. Our data indicate that the differentially expressed proteins participate in a variety of cellular processes and enhance our understanding of the mechanisms of drug resistance in ovarian cancer cells.  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