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 release of cytochrome c from mitochondria, which leads to activation of the intrinsic apoptotic pathway, is regulated by interactions of Bax and Bak with antiapoptotic Bcl-2 family members. The factors that regulate these interactions are, at the present time, incompletely understood. Recent studies showing preferences in binding between synthetic Bcl-2 homology domain 3 and antiapoptotic Bcl-2 family members in vitro have suggested that the antiapoptotic proteins Mcl-1 and Bcl-xL, but not Bcl-2, restrain proapoptotic Bak from inducing mitochondrial membrane permeabilization and apoptosis. Here we show that Bak protein has a much higher affinity than the 26-amino acid Bak Bcl-2 homology domain 3 for Bcl-2, that some naturally occurring Bcl-2 allelic variants have an affinity for full-length Bak that is only 3-fold lower than that of Mcl-1, and that endogenous levels of these Bcl-2 variants (which are as much as 40-fold more abundant than Mcl-1) restrain part of the Bak in intact lymphoid cells. In addition, we demonstrate that Bcl-2 variants can, depending on their affinity for Bak, substitute for Mcl-1 in protecting cells. Thus, the ability of Bcl-2 to protect cells from activated Bak depends on two important contextual variables, the identity of the Bcl-2 present and the amount expressed.The release of cytochrome c from mitochondria, which leads to activation of the intrinsic apoptotic pathway, is regulated by Bcl-2 family members (15). This group of proteins consists of three subgroups: Bax and Bak, which oligomerize upon death stimulation to form a putative pore in the outer mitochondrial membrane, thereby allowing efflux of cytochrome c and other mitochondrial intermembrane space components; Bcl-2, Bcl-xL, Mcl-1, and other antiapoptotic homologs, which antagonize the effects of Bax and Bak; and BH3-only proteins2 such as Bim, Bid, and Puma, which are proapoptotic Bcl-2 family members that share only limited homology with the other two groups in a single 15-amino acid domain (the BH3 domain, see Ref. 6). Although it is clear that BH3-only proteins serve as molecular sensors of various stresses and, when activated, trigger apoptosis (3, 611), the mechanism by which they do so remains incompletely understood. One current model suggests that BH3-only proteins trigger apoptosis solely by binding and neutralizing antiapoptotic Bcl-2 family members, thereby causing them to release the activated Bax and Bak that are bound (reviewed in Refs. 9 and 10; see also Refs. 12 and 13), whereas another current model suggests that certain BH3-only proteins also directly bind to and activate Bax (reviewed in Ref. 3; see also Refs. 1417). Whichever model turns out to be correct, both models agree that certain antiapoptotic Bcl-2 family members can inhibit apoptosis, at least in part, by binding and neutralizing activated Bax and Bak before they permeabilize the outer mitochondrial membrane (13, 18, 19).Much of the information about the interactions between pro- and antiapoptotic Bcl-2 family members has been derived from the study of synthetic peptides corresponding to BH3 domains. In particular, these synthetic peptides have been utilized as surrogates for the full-length proapoptotic proteins during structure determinations (2022) as well as in functional studies exploring the effect of purified BH3 domains on isolated mitochondria (14, 23) and on Bax-mediated permeabilization of lipid vesicles (15).Recent studies using these same peptides have suggested that interactions of the BH3 domains of Bax, Bak, and the BH3-only proteins with the “BH3 receptors” of the antiapoptotic Bcl-2 family members are not all equivalent. Surface plasmon resonance, a technique that is widely used to examine the interactions of biomolecules under cell-free conditions (2426), has demonstrated that synthetic BH3 peptides of some BH3-only family members show striking preferences, with the Bad BH3 peptide binding to Bcl-2 and Bcl-xL but not Mcl-1, and the Noxa BH3 peptide binding to Mcl-1 but not Bcl-2 or Bcl-xL (27). Likewise, the Bak BH3 peptide exhibits selectivity, with high affinity for Bcl-xL and Mcl-1 but not Bcl-2 (12). The latter results have led to a model in which Bcl-xL and Mcl-1 restrain Bak and inhibit Bak-dependent apoptosis, whereas Bcl-2 does not (10).Because the Bak protein contains multiple recognizable domains in addition to its BH3 motif (28, 29), we compared the binding of Bak BH3 peptide and Bak protein to Bcl-2. Surface plasmon resonance demonstrated that Bcl-2 binds Bak protein with much higher affinity than the Bak 26-mer BH3 peptide. Further experiments demonstrated that the KD for Bak differs among naturally occurring Bcl-2 sequence variants but is only 3-fold higher than that of Mcl-1 in some cases. In light of previous reports that Bcl-2 overexpression contributes to neoplastic transformation (3033) and drug resistance (3436) in lymphoid cells, we also examined Bcl-2 expression and Bak binding in a panel of neoplastic lymphoid cell lines. Results of these experiments demonstrated that Bcl-2 expression varies among different lymphoid cell lines but is up to 40-fold more abundant than Mcl-1. In lymphoid cell lines with abundant Bcl-2, Bak is detected in Bcl-2 as well as Mcl-1 immunoprecipitates; and Bak-dependent apoptosis induced by Mcl-1 down-regulation can be prevented by Bcl-2 overexpression. Collectively, these observations shed new light on the role of Bcl-2 in binding and neutralizing Bak.  相似文献   

2.
3.
The mechanisms of free fatty acid-induced lipoapoptosis are incompletely understood. Here we demonstrate that Mcl-1, an anti-apoptotic member of the Bcl-2 family, was rapidly degraded in hepatocytes in response to palmitate and stearate by a proteasome-dependent pathway. Overexpression of a ubiquitin-resistant Mcl-1 mutant in Huh-7 cells attenuated palmitate-mediated Mcl-1 loss and lipoapoptosis; conversely, short hairpin RNA-targeted knockdown of Mcl-1 sensitized these cells to lipoapoptosis. Palmitate-induced Mcl-1 degradation was attenuated by the novel protein kinase C (PKC) inhibitor rottlerin. Of the two human novel PKC isozymes, PKCδ and PKCθ, only activation of PKCθ was observed by phospho-immunoblot analysis. As compared with Jurkat cells, a smaller PKCθ polypeptide and mRNA were expressed in hepatocytes consistent with an alternative splice variant. Short hairpin RNA-mediated knockdown of PKCθ reduced Mcl-1 degradation and lipoapoptosis. Likewise, genetic deletion of Pkcθ also attenuated Mcl-1 degradation and cytotoxicity by palmitate in primary hepatocytes. During treatment with palmitate, rottlerin inhibited phosphorylation of Mcl-1 at Ser159, a phosphorylation site previously implicated in Mcl-1 turnover. Consistent with these results, an Mcl-1 S159A mutant was resistant to degradation and improved cell survival during palmitate treatment. Collectively, these results implicate PKCθ-dependent destabilization of Mcl-1 as a mechanism contributing to hepatocyte lipoapoptosis.Current evidence suggests that hepatic steatosis is present in up to 30% of the American population (1). A subset of these individuals develop severe hepatic lipotoxicity, a syndrome referred to as NASH2 (2), which can progress to cirrhosis and its chronic sequela (3, 4). A major risk factor for hepatic lipotoxicity is insulin resistance (57), resulting in excessive lipolysis within peripheral adipose tissue with release of high levels of free fatty acids (FFA) to the circulation. Circulating FFA are taken up by the liver via fatty acid transporter 5 and CD36 (810), and the bulk of hepatic neutral fat is derived from re-esterification of circulating FFA (8). Current concepts indicate that FFA, and not their esterified product (triglyceride), mediate hepatic lipotoxicity (11, 12). Elevated serum FFA correlate with liver disease severity (1315), and therapies that enhance insulin sensitivity ameliorate hepatic lipotoxicity, in part, by decreasing plasma FFA (16). Hepatic FFA also accumulate in experimental steatohepatitis, further supporting a role for these nutrients in hepatic lipotoxicity (17). Saturated FFA are more strongly implicated in hepatic lipotoxicity than unsaturated FFA (18, 19). Saturated FFA induce hepatocyte apoptosis (20, 21), a cardinal feature of nonalcoholic fatty liver disease (22), and serum biomarkers of apoptosis are useful for identifying hepatic lipotoxicity (23). Thus, FFA-mediated lipotoxicity occurs, in part, by apoptosis.Apoptosis is regulated by members of the Bcl-2 protein family (24). These proteins can be categorized into three subsets as follows: the guardians or anti-apoptotic members of this family, which include Bcl-2, A1, Mcl-1, Bcl-xL, and Bcl-w; the multidomain executioners or proapoptotic members of this family, which include Bax and Bak; and the messengers or biosensors of cell death, which share only the third Bcl-2 homology domain and are referred to as BH3-only proteins. This last group of proteins includes Bid, Bim, Bmf, Puma, Noxa, Hrk, Bad, and Bik. We have previously reported that cytotoxic FFA induce Bim expression by a FoxO3a-dependent mechanism that contributes, in part, to lipoapoptosis by activating Bax (20, 21). However, Bax activation can be held in check by anti-apoptotic members of the Bcl-2 family suggesting their function may also be dysregulated during FFA-mediated cytotoxicity.Bcl-2 is not expressed in hepatocytes at the protein level (25), whereas Bcl-w and Bfl-1/A1 knock-out mice have no liver phenotype (2628). However, both potent anti-apoptotic proteins Bcl-xL and Mcl-1 are expressed by hepatocytes and exhibit a liver phenotype in knock-out mice (29, 30), whereas up-regulation of Mcl-1 renders hepatocytes resistant to apoptosis (3133). It has also been posited that cellular elimination of Mcl-1 is a critical step in certain proapoptotic cascades (34, 35). Mcl-1 is unique among Bcl-2 proteins in that it has a short half-life, 30–120 min in most cell types, due to the presence of two sequences rich in proline, glutamic acid, serine, and threonine, which target the protein for rapid degradation by the proteasome (36). Proteasomal degradation of Mcl-1 is promoted by ubiquitination, which in turn is regulated by various kinase cascades (36). Despite its potential importance, a role for Mcl-1 in regulating hepatocyte FFA-mediated lipoapoptosis remains unexplored.Given that FFA induce insulin resistance (37), the kinases potentially regulating lipoapoptosis are likely those also identified in insulin resistance syndromes, especially the novel PKC isoforms PKCδ and PKCθ (38). The novel PKC isoforms are activated by diacylglycerol, which rises in the presence of FFA (3941), and diacylglycerol levels are significantly increased in NASH (42). A role for PKCδ in apoptosis has not been described. PKCθ has recently been shown to be activated by endoplasmic reticulum stress in liver cells (43) and lipids in vivo (44, 45). Furthermore, PKCθ has also been implicated in apoptosis of Jurkat cells, neuroblastoma cells, and myeloid leukemia cells (46, 47). However, neither its role in mediating lipoapoptosis nor modulating levels/activity of Bcl-2 proteins has been examined.This study addresses the role of Mcl-1 and PKCθ in FFA-induced lipoapoptosis. We identify a pathway that involves PKCθ-dependent proteasomal degradation of Mcl-1. Using inhibitors of various steps along this pathway, along with Mcl-1 mutants that are resistant to proteasomal degradation or Ser159 phosphorylation, our studies implicate Mcl-1 degradation via a PKCθ-dependent process as a critical step in lipoapoptosis.  相似文献   

4.
The molecular chaperone Hsp27 exists as a distribution of large oligomers that are disassembled by phosphorylation at Ser-15, -78, and -82. It is controversial whether the unphosphorylated Hsp27 or the widely used triple Ser-to-Asp phospho-mimic mutant is the more active molecular chaperone in vitro. This question was investigated here by correlating chaperone activity, as measured by the aggregation of reduced insulin or α-lactalbumin, with Hsp27 self-association as monitored by analytical ultracentrifugation. Furthermore, because the phospho-mimic is generally assumed to reproduce the phosphorylated molecule, the size and chaperone activity of phosphorylated Hsp27 were compared with that of the phospho-mimic. Hsp27 was triply phosphorylated by MAPKAP-2 kinase, and phosphorylation was tracked by urea-PAGE. An increasing degree of suppression of insulin or α-lactalbumin aggregation correlated with a decreasing Hsp27 self-association, which was the least for phosphorylated Hsp27 followed by the mimic followed by the unphosphorylated protein. It was also found that Hsp27 added to pre-aggregated insulin did not reverse aggregation but did inhibit these aggregates from assembling into even larger aggregates. This chaperone activity appears to be independent of Hsp27 phosphorylation. In conclusion, the most active chaperone of insulin and α-lactalbumin was the Hsp27 (elongated) dimer, the smallest Hsp27 subunit observed under physiological conditions. Next, the Hsp27 phospho-mimic is only a partial mimic of phosphorylated Hsp27, both in self-association and in chaperone function. Finally, the efficient inhibition of insulin aggregation by Hsp27 dimer led to the proposal of two models for this chaperone activity.Oligomeric heat shock protein 27 (Hsp27)2 is a ubiquitous mammalian protein with a variety of functions in health and disease (18). These functions include ATP-independent chaperone activity in response to environmental stress, e.g. heat shock and oxidative stress, control of apoptosis, and regulation of actin cytoskeleton dynamics. Hsp27 is a member of the α-crystallin small heat shock protein family of which αB-crystallin is the archetype. These proteins are characterized by an α-crystallin domain of 80–90 residues consisting of roughly eight β-strands that form an intermolecular β-sheet interaction interface within a dimer, the basic building subunit of the oligomer (2, 4, 911).Hsp27 is in equilibrium between high molecular weight oligomers and much lower molecular weight multimers. It has been reported that unphosphorylated Hsp27 includes predominantly a distribution of high molecular species ranging in size from 12-mer to 35-mer (1219). Phosphorylation of Hsp27 at serines 15, 78, and 82 by the p38-activated MAPKAP-2 kinase (2022) or the use of the triple Ser-to-Asp phospho-mimic results in a major shift in the equilibrium toward much smaller multimers (23) and in an alteration of its function (1, 3, 6, 7, 24, 25). The size distribution of the smaller species has been reported to be between monomer and tetramer (1216, 18, 19).Small heat shock proteins, including Hsp27, behave as ATP-independent molecular chaperones during cellular heat shock. They bind partially unfolded proteins and prevent their aggregation until the proteins can be refolded by larger ATP-dependent chaperones or are digested (7, 8, 26). This function includes the up-regulation and/or phosphorylation of Hsp27.It is not entirely clear what the role of Hsp27 size and phosphorylation state plays in its heat shock function because there are conflicting results in the literature. Some in vitro studies concluded that the unphosphorylated oligomeric Hsp27 (or the murine isoform Hsp25) protects proteins against aggregation better than does the phosphorylation mimic (13, 19, 27), whereas others found no difference (16, 28, 29), and still other studies found that the mimic protects better than does the unphosphorylated wild type (27, 30, 31). In-cell studies found that phosphorylation of Hsp27 was essential for thermo-protection of actin filaments (32), and the Hsp27 phosphorylation mimic decreased inclusion body formation better than did unphosphorylated Hsp27 (33). This study was undertaken to investigate the molecular chaperone function of Hsp27 by correlating chaperone activity with Hsp27 size and by comparing fully phosphorylated Hsp27 with its phospho-mimic.  相似文献   

5.
The Dbf4-Cdc7 kinase (DDK) is required for the activation of the origins of replication, and DDK phosphorylates Mcm2 in vitro. We find that budding yeast Cdc7 alone exists in solution as a weakly active multimer. Dbf4 forms a likely heterodimer with Cdc7, and this species phosphorylates Mcm2 with substantially higher specific activity. Dbf4 alone binds tightly to Mcm2, whereas Cdc7 alone binds weakly to Mcm2, suggesting that Dbf4 recruits Cdc7 to phosphorylate Mcm2. DDK phosphorylates two serine residues of Mcm2 near the N terminus of the protein, Ser-164 and Ser-170. Expression of mcm2-S170A is lethal to yeast cells that lack endogenous MCM2 (mcm2Δ); however, this lethality is rescued in cells harboring the DDK bypass mutant mcm5-bob1. We conclude that DDK phosphorylation of Mcm2 is required for cell growth.The Cdc7 protein kinase is required throughout the yeast S phase to activate origins (1, 2). The S phase cyclin-dependent kinase also activates yeast origins of replication (35). It has been proposed that Dbf4 activates Cdc7 kinase in S phase, and that Dbf4 interaction with Cdc7 is essential for Cdc7 kinase activity (6). However, it is not known how Dbf4-Cdc7 (DDK)2 acts during S phase to trigger the initiation of DNA replication. DDK has homologs in other eukaryotic species, and the role of Cdc7 in activation of replication origins during S phase may be conserved (710).The Mcm2-7 complex functions with Cdc45 and GINS to unwind DNA at a replication fork (1115). A mutation of MCM5 (mcm5-bob1) bypasses the cellular requirements for DBF4 and CDC7 (16), suggesting a critical physiologic interaction between Dbf4-Cdc7 and Mcm proteins. DDK phosphorylates Mcm2 in vitro with proteins purified from budding yeast (17, 18) or human cells (19). Furthermore, there are mutants of MCM2 that show synthetic lethality with DBF4 mutants (6, 17), suggesting a biologically relevant interaction between DBF4 and MCM2. Nevertheless, the physiologic role of DDK phosphorylation of Mcm2 is a matter of dispute. In human cells, replacement of MCM2 DDK-phosphoacceptor residues with alanines inhibits DNA replication, suggesting that Dbf4-Cdc7 phosphorylation of Mcm2 in humans is important for DNA replication (20). In contrast, mutation of putative DDK phosphorylation sites at the N terminus of Schizosaccharomyces pombe Mcm2 results in viable cells, suggesting that phosphorylation of S. pombe Mcm2 by DDK is not critical for cell growth (10).In budding yeast, Cdc7 is present at high levels in G1 and S phase, whereas Dbf4 levels peak in S phase (18, 21, 22). Furthermore, budding yeast DDK binds to chromatin during S phase (6), and it has been shown that Dbf4 is required for Cdc7 binding to chromatin in budding yeast (23, 24), fission yeast (25), and Xenopus (9). Human and fission yeast Cdc7 are inert on their own (7, 8), but Dbf4-Cdc7 is active in phosphorylating Mcm proteins in budding yeast (6, 26), fission yeast (7), and human (8, 10). Based on these data, it has been proposed that Dbf4 activates Cdc7 kinase in S phase and that Dbf4 interaction with Cdc7 is essential for Cdc7 kinase activity (6, 9, 18, 2124). However, a mechanistic analysis of how Dbf4 activates Cdc7 has not yet been accomplished. For example, the multimeric state of the active Dbf4-Cdc7 complex is currently disputed. A heterodimer of fission yeast Cdc7 (Hsk1) in complex with fission yeast Dbf4 (Dfp1) can phosphorylate Mcm2 (7). However, in budding yeast, oligomers of Cdc7 exist in the cell (27), and Dbf4-Cdc7 exists as oligomers of 180 and 300 kDa (27).DDK phosphorylates the N termini of human Mcm2 (19, 20, 28), human Mcm4 (10), budding yeast Mcm4 (26), and fission yeast Mcm6 (10). Although the sequences of the Mcm N termini are poorly conserved, the DDK sites identified in each study have neighboring acidic residues. The residues of budding yeast Mcm2 that are phosphorylated by DDK have not yet been identified.In this study, we find that budding yeast Cdc7 is weakly active as a multimer in phosphorylating Mcm2. However, a low molecular weight form of Dbf4-Cdc7, likely a heterodimer, has a higher specific activity for phosphorylation of Mcm2. Dbf4 or DDK, but not Cdc7, binds tightly to Mcm2, suggesting that Dbf4 recruits Cdc7 to Mcm2. DDK phosphorylates two serine residues of Mcm2, Ser-164 and Ser-170, in an acidic region of the protein. Mutation of Ser-170 is lethal to yeast cells, but this phenotype is rescued by the DDK bypass mutant mcm5-bob1. We conclude that DDK phosphorylation of Ser-170 of Mcm2 is required for budding yeast growth.  相似文献   

6.
High accuracy mass spectrometry has proven to be a powerful technology for the large scale identification of serine/threonine/tyrosine phosphorylation in the living cell. However, despite many described phosphoproteomes, there has been no comparative study of the extent of phosphorylation and its evolutionary conservation in all domains of life. Here we analyze the results of phosphoproteomics studies performed with the same technology in a diverse set of organisms. For the most ancient organisms, the prokaryotes, only a few hundred proteins have been found to be phosphorylated. Applying the same technology to eukaryotic species resulted in the detection of thousands of phosphorylation events. Evolutionary analysis shows that prokaryotic phosphoproteins are preferentially conserved in all living organisms, whereas-site specific phosphorylation is not. Eukaryotic phosphosites are generally more conserved than their non-phosphorylated counterparts (with similar structural constraints) throughout the eukaryotic domain. Yeast and Caenorhabditis elegans are two exceptions, indicating that the majority of phosphorylation events evolved after the divergence of higher eukaryotes from yeast and reflecting the unusually large number of nematode-specific kinases. Mitochondria present an interesting intermediate link between the prokaryotic and eukaryotic domains. Applying the same technology to this organelle yielded 174 phosphorylation sites mapped to 74 proteins. Thus, the mitochondrial phosphoproteome is similarly sparse as the prokaryotic phosphoproteomes. As expected from the endosymbiotic theory, phosphorylated as well as non-phosphorylated mitochondrial proteins are significantly conserved in prokaryotes. However, mitochondrial phosphorylation sites are not conserved throughout prokaryotes, consistent with the notion that serine/threonine phosphorylation in prokaryotes occurred relatively recently in evolution. Thus, the phosphoproteome reflects major events in the evolution of life.Reversible protein phosphorylation on serines, threonines, and tyrosines plays a crucial role in regulating processes in all living organisms ranging from prokaryotes to eukaryotes (1). Traditionally, phosphorylation has been detected in single, purified proteins using in vitro assays. Recent advances in mass spectrometry (MS)-based proteomics now allow the identification of in vivo phosphorylation sites with high accuracy (27). On-line databases such as PhosphoSite (8), Phospho.ELM (9), and PHOSIDA1 (10) have collected and organized thousands of identified phosphosites. These databases as well as dedicated analysis environments such as NetworKIN (11, 12) offer and use contextual information including structural features, potential kinases, and conservation. They constitute resources that should allow the derivation of general patterns for phosphorylation events. Specifically, the recent availability of data for archaeal, prokaryotic, and diverse eukaryotic phosphoproteomes in these databases should enable investigation of the evolutionary history of this post-translational modification.Prokaryotes have two separate classes of phosphorylation events. Apart from the canonical histidine/aspartate phosphorylation, which has been studied for decades, serine/threonine/tyrosine phosphorylation is also present and has recently become amenable to analysis by MS (13). Bacterial phosphoproteins are involved in protein synthesis, carbohydrate metabolism, and the phosphoenolpyruvate-dependent phosphotransferase system. Recent phosphoproteomics studies of Bacillus subtilis, Escherichia coli, and Lactococcus lactis described around 100 phosphorylation sites on serine, threonine, and tyrosine in each of these species (1315). Bacterial phosphorylation sites can change in response to environmental conditions (16).Interestingly, even archaea have serine/threonine and tyrosine phosphorylation. A recent study of Halobacterium salinarum described 75 serine/threonine/tyrosine phosphorylation sites on 62 proteins involved in a wide range of cellular processes including a variety of metabolic pathways (17).Although only a few hundred phosphorylation events have been found in prokaryotic species, similar experimental conditions and effort have yielded the detection of thousands of phosphorylation events in eukaryotes ranging from yeast to human (7, 1821). Before the advent of large scale phosphoproteomics, serine/threonine/tyrosine phosphorylation has been estimated to affect one-third of all proteins (22). Recent large scale phosphoproteomics studies now suggest that more than half of all eukaryotic proteins are phosphorylated (23).A key event in evolution was the endosymbiosis of prokaryotes that enabled the development of a much more complex type of life, the eukaryotic cell. Analyses of mitochondrial genes suggest that the α-proteobacterium Rickettsia prowazekii is the endosymbiotic precursor leading to modern mitochondria (24). Almost all of the mitochondrial genes have migrated to the nuclear genome during subsequent evolution, and it is predicted that 10–15% of eukaryotic nuclear genes of organisms encode mitochondrial proteins (25).Thus, mitochondria with their unique evolutionary position between prokaryotes and eukaryotes form an interesting link for the evolutionary analysis of phosphorylation. Several studies investigated the mitochondrial phosphoproteome in different organisms using gel electrophoresis or specific enrichment methods coupled with mass spectrometry (2628). Those studies established potential mitochondrial phosphoproteins. Three large scale studies based on affinity enrichment of phosphopeptides and mass spectrometry obtained direct experimental evidence of phosphorylation sites in mitochondria. Lee et al. (29) used a combination of different peptide enrichment strategies and found 80 phosphorylation sites of 48 different proteins from mouse liver. Very recently, a study by Deng et al. (30) characterized the murine cardiac mitochondrial mouse phosphoproteome, covering 236 phosphosites on 181 proteins. Investigating yeast, Reinders et al. (31) assigned 84 phosphorylation sites in 62 proteins.To enable comparative analysis of phosphoproteomes between all domains of life and mitochondria, here we experimentally determined a high accuracy mitochondrial mouse phosphoproteome based on technology conditions similar to those applied to the identification of prokaryotic and eukaryotic phosphoproteomes. We then performed a detailed evolutionary study of the conservation of the identified phosphoproteins and phosphorylation sites in prokaryotes and in eukaryotes. This allowed an initial comparison of the phosphoproteomes of prokaryotes, mitochondria, and eukaryotes.  相似文献   

7.
Mitochondria play a central role in energy metabolism and cellular survival, and consequently mitochondrial dysfunction is associated with a number of human pathologies. Reversible protein phosphorylation emerges as a central mechanism in the regulation of several mitochondrial processes. In skeletal muscle, mitochondrial dysfunction is linked to insulin resistance in humans with obesity and type 2 diabetes. We performed a phosphoproteomics study of functional mitochondria isolated from human muscle biopsies with the aim to obtain a comprehensive overview of mitochondrial phosphoproteins. Combining an efficient mitochondrial isolation protocol with several different phosphopeptide enrichment techniques and LC-MS/MS, we identified 155 distinct phosphorylation sites in 77 mitochondrial phosphoproteins, including 116 phosphoserine, 23 phosphothreonine, and 16 phosphotyrosine residues. The relatively high number of phosphotyrosine residues suggests an important role for tyrosine phosphorylation in mitochondrial signaling. Many of the mitochondrial phosphoproteins are involved in oxidative phosphorylation, tricarboxylic acid cycle, and lipid metabolism, i.e. processes proposed to be involved in insulin resistance. We also assigned phosphorylation sites in mitochondrial proteins involved in amino acid degradation, importers and transporters, calcium homeostasis, and apoptosis. Bioinformatics analysis of kinase motifs revealed that many of these mitochondrial phosphoproteins are substrates for protein kinase A, protein kinase C, casein kinase II, and DNA-dependent protein kinase. Our results demonstrate the feasibility of performing phosphoproteome analysis of organelles isolated from human tissue and provide novel targets for functional studies of reversible phosphorylation in mitochondria. Future comparative phosphoproteome analysis of mitochondria from healthy and diseased individuals will provide insights into the role of abnormal phosphorylation in pathologies, such as type 2 diabetes.Mitochondria are the primary energy-generating systems in eukaryotes. They play a crucial role in oxidative metabolism, including carbohydrate metabolism, fatty acid oxidation, and urea cycle, as well as in calcium signaling and apoptosis (1, 2). Mitochondrial dysfunction is centrally involved in a number of human pathologies, such as type 2 diabetes, Parkinson disease, and cancer (3). The most prevalent form of cellular protein post-translational modifications (PTMs),1 reversible phosphorylation (46), is emerging as a central mechanism in the regulation of mitochondrial functions (7, 8). The steadily increasing numbers of reported mitochondrial kinases, phosphatases, and phosphoproteins imply an important role of protein phosphorylation in different mitochondrial processes (911).Mass spectrometry (MS)-based proteome analysis is a powerful tool for global profiling of proteins and their PTMs, including protein phosphorylation (12, 13). A variety of proteomics techniques have been developed for specific enrichment of phosphorylated proteins and peptides and for phosphopeptide-specific data acquisition techniques at the MS level (14). Enrichment methods based on affinity chromatography, such as titanium dioxide (TiO2) (1517), zwitterionic hydrophilic interaction chromatography (ZIC-HILIC) (18), immobilized metal affinity chromatography (IMAC) (19, 20), and ion exchange chromatography (strong anion exchange and strong cation exchange) (21, 22), have shown high efficiencies for enrichment of phosphopeptides (14). Recently, we demonstrated that calcium phosphate precipitation (CPP) is highly effective for enriching phosphopeptides (23). It is now generally accepted that no single method is comprehensive, but combinations of different enrichment methods produce distinct overlapping phosphopeptide data sets to enhance the overall results in phosphoproteome analysis (24, 25). Phosphopeptide sequencing by mass spectrometry has seen tremendous advances during the last decade (26). For example, MS/MS product ion scanning, multistage activation, and precursor ion scanning are effective methods for identifying serine (Ser), threonine (Thr), and tyrosine (Tyr) phosphorylated peptides (14, 26).A “complete” mammalian mitochondrial proteome was reported by Mootha and co-workers (27) and included 1098 proteins. The mitochondrial phosphoproteome has been characterized in a series of studies, including yeast, mouse and rat liver, porcine heart, and plants (19, 2831). To date, the largest data set by Deng et al. (30) identified 228 different phosphoproteins and 447 phosphorylation sites in rat liver mitochondria. However, the in vivo phosphoproteome of human mitochondria has not been determined. A comprehensive mitochondrial phosphoproteome is warranted for further elucidation of the largely unknown mechanisms by which protein phosphorylation modulates diverse mitochondrial functions.The percutaneous muscle biopsy technique is an important tool in the diagnosis and management of human muscle disorders and has been widely used to investigate metabolism and various cellular and molecular processes in normal and abnormal human muscle, in particular the molecular mechanism underlying insulin resistance in obesity and type 2 diabetes (32). Skeletal muscle is rich in mitochondria and hence a good source for a comprehensive proteomics and functional analysis of mitochondria (32, 33).The major aim of the present study was to obtain a comprehensive overview of site-specific phosphorylation of mitochondrial proteins in functionally intact mitochondria isolated from human skeletal muscle. Combining an efficient protocol for isolation of skeletal muscle mitochondria with several different state-of-the-art phosphopeptide enrichment methods and high performance LC-MS/MS, we identified 155 distinct phosphorylation sites in 77 mitochondrial phosphoproteins, many of which have not been reported before. We characterized this mitochondrial phosphoproteome by using bioinformatics tools to classify functional groups and functions, including kinase substrate motifs.  相似文献   

8.
Phosphorylation is a major post-translational modification widely used in the regulation of many cellular processes. Cyclin-dependent kinase 5 (Cdk5) is a proline-directed serine/threonine kinase activated by activation subunit p35. Cdk5-p35 regulates various neuronal activities such as neuronal migration, spine formation, synaptic activity, and cell death. The kinase activity of Cdk5 is regulated by proteolysis of p35: proteasomal degradation causes down-regulation of Cdk5, whereas cleavage of p35 by calpain causes overactivation of Cdk5. Phosphorylation of p35 determines the proteolytic pathway. We have previously identified Ser8 and Thr138 as major phosphorylation sites using metabolic labeling of cultured cells followed by two-dimensional phosphopeptide mapping and phosphospecific antibodies. However, these approaches cannot determine the extent of p35 phosphorylation in vivo. Here we report the use of Phos-tag SDS-PAGE to reveal the phosphorylation states of p35 in neuronal culture and brain. Using Phos-tag acrylamide, the electrophoretic mobility of phosphorylated p35 was delayed because it is trapped at Phos-tag sites. We found a novel phosphorylation site at Ser91, which was phosphorylated by Ca2+-calmodulin-dependent protein kinase II in vitro. We constructed phosphorylation-dependent banding profiles of p35 and Ala substitution mutants at phosphorylation sites co-expressed with Cdk5 in COS-7 cells. Using the standard banding profiles, we assigned respective bands of endogenous p35 with combinations of phosphorylation states and quantified Ser8, Ser91, and Thr138 phosphorylation. The highest level of p35 phosphorylation was observed in embryonic brain; Ser8 was phosphorylated in all p35 molecules, whereas Ser91 was phosphorylated in 60% and Thr138 was phosphorylated in ∼12% of p35 molecules. These are the first quantitative and site-specific measurements of phosphorylation of p35, demonstrating the usefulness of Phos-tag SDS-PAGE for analysis of phosphorylation states of in vivo proteins.Phosphorylation is a major post-translational modification of proteins, modulating a variety of cellular functions (1, 2). Because most phosphorylation occurs in a highly site-specific manner, identification of phosphorylation sites has been a subject of intense investigation. Several analytical methods have been utilized to identify phosphorylation sites, including mass spectrometry, amino acid sequencing, and radioisotope phosphate labeling of proteins with mutation(s) at putative phosphorylation site(s) (3, 4). Phosphorylation site-specific antibodies are frequently used to detect phosphorylation at target sites (5, 6). Many phosphospecific antibodies are now commercially available. These phosphospecific antibodies are convenient and useful tools for examining site-specific phosphorylation both in vivo and in vitro. However, they are not appropriate for estimating quantitative ratios of phosphorylation states. Electrophoretic mobility shift on SDS-PAGE is also often used to observe phosphorylation (710), but this method is not always applied to site-specific phosphorylation.Phos-tag is a newly developed dinuclear metal complex that can be used to provide phosphate-binding sites when conjugated to analytical materials such as acrylamide and biotin (11). In SDS-PAGE using Phos-tag acrylamide, phosphorylated proteins are trapped by the Phos-tag sites, delaying their migration and thus separating them from unphosphorylated proteins. Subsequent immunoblot analysis with phosphorylation-independent antibodies reveals both the phosphorylated and unphosphorylated bands. Because the migration of the phosphorylated proteins is greatly delayed compared with migration in Laemmli SDS-PAGE, it is easy to identify the phosphorylated proteins from observed positions on blots. In the past 3 years, this method has been used to detect phosphorylation states for many proteins such as ERK1/2, cdc37, myosin light chain, eIF2α, protein kinase D, β-casein, SIRT7, and dysbindin-1 (1221).Cyclin-dependent kinase 5 (Cdk5)1 is a proline-directed serine/threonine kinase that is expressed predominantly in postmitotic neurons and regulates various neuronal events such as neuronal migration, spine formation, synaptic activity, and cell death (2224). Cdk5 is activated by binding to activation subunit p35 and inactivated by proteasomal degradation of p35 (25). In addition, Cdk5 activity is deregulated by cleavage of p35 to p25 with calpain, resulting in abnormal activation and ultimately causing neuronal cell death (2629). Proteolysis of p35, either by proteasomal degradation or cleavage by calpain, is regulated by phosphorylation of p35 by Cdk5 (3033). Therefore, phosphorylation of p35 is essential for proper regulation of Cdk5 activity and function. We previously identified Ser8 and Thr138 as major p35 phosphorylation sites (33). We also showed that phosphorylation of p35 decreased during brain development and proposed its relationship to age-dependent vulnerability of neurons to stress stimuli (32). Thus, to understand the in vivo regulation of Cdk5 activity, it is critical to analyze the phosphorylation states of p35 in brain. However, there is no convenient method to analyze the precise in vivo phosphorylation status of the endogenous proteins.In this study, we applied the Phos-tag SDS-PAGE method to analyze the phosphorylation states of p35 in vivo and in cultured neurons. We constructed standard band profiles of phosphorylated p35 by Phos-tag SDS-PAGE using Ala mutants at Ser8 and/or Thr138. From these experiments, we observed an unidentified in vivo phosphorylation site at Ser91. We quantified the phosphorylation at each site in cultured neurons and brain, providing the first quantitative estimate of the in vivo phosphorylation states of p35. We discuss the usefulness of Phos-tag SDS-PAGE to analyze the in vivo phosphorylation states of proteins.  相似文献   

9.
10.
Emerging evidence suggests that the resistance of cancer stem cells (CSC) to many conventional therapies is one of the major limiting factors of cancer therapy efficacy. Identification of mechanisms responsible for survival and self-renewal of CSC will help design new therapeutic strategies that target and eliminate both differentiated cancer cells and CSC. Here we demonstrated the potential role of proapoptotic protein BAD in the biology of CSC in melanoma, prostate and breast cancers. We enriched CD44+/CD24 cells (CSC) by tumorosphere formation and purified this population by FACS. Both spheres and CSC exhibited increased potential for proliferation, migration, invasion, sphere formation, anchorage-independent growth, as well as upregulation of several stem cell-associated markers. We showed that the phosphorylation of BAD is essential for the survival of CSC. Conversely, ectopic expression of a phosphorylation-deficient mutant BAD induced apoptosis in CSC. This effect was enhanced by treatment with a BH3-mimetic, ABT-737. Both pharmacological agents that inhibit survival kinases and growth factors that are involved in drug resistance delivered their respective cytotoxic and protective effects by modulating the BAD phosphorylation in CSC. Furthermore, the frequency and self-renewal capacity of CSC was significantly reduced by knocking down the BAD expression. Consistent with our in vitro results, significant phosphorylation of BAD was found in CD44+ CSC of 83% breast tumor specimens. In addition, we also identified a positive correlation between BAD expression and disease stage in prostate cancer, suggesting a role of BAD in tumor advancement. Our studies unveil the role of BAD in the survival and self-renewal of CSC and propose BAD not only as an attractive target for cancer therapy but also as a marker of tumor progression.Although tumors initially respond positively to anti-cancer agents, several cancers, despite the best care and significant improvements in treatment, recur and progress to advanced stages of the disease. The mechanisms underlying this recurrence and metastasis are not clearly understood. Over the past decade, substantial evidence supported the cancer stem cell (CSC) hypothesis as a viable explanation for the initiation, progression and recurrence of cancer. According to this hypothesis, each tumor harbors a small subpopulation of specialized cells among cellular heterogeneity, known as CSC. These cells exhibit self-renewal property that drives tumorigenesis and plasticity to differentiate into multiple cell types contributing to tumor cellular heterogeneity. Support for this hypothesis came from the studies by Lapidot et al. who identified tumor-initiating cells in acute myeloid leukemia.1, 2 Subsequently, CSCs have been identified in several cancers.3, 4, 5, 6, 7, 8, 9, 10Accumulating evidence suggests that current cancer therapies can only shrink tumors as they target and kill the differentiated cancer (DC) cells, but are unable to target the rare CSC population.11, 12 Thus, despite a wealth of information on DC cells, the active survival and self-renewal pathways in CSCs have not been characterized thoroughly. An understanding of the molecular mechanisms involved in the survival, self-renewal and resistance of CSCs to current therapeutic regimens is of immense clinical interest. This information will help in developing novel strategies for more effective treatments for cancer.Most anti-cancer drugs exert their effects through triggering the apoptotic pathways. However, malignant cancer cells can escape apoptosis by altering the expression level of proapoptotic and antiapoptotic BCL-2 family members. Considering the potential role of BCL-2 family members in tumorigenesis and cancer cell survival, their role in CSC biology has been increasingly studied.13, 14 BAD (BCL2-antagonist of cell death) is a member of the BH3-only BCL-2 family protein that when dephosphorylated promotes apoptosis by heterodimerizing with the antiapoptotic proteins BCL-XL and BCL-2.15 The cytotoxic effects of BAD are controlled by mechanisms that regulate its phosphorylation on at least two distinct serine residues, S112 and S136.16, 17, 18 Previously, we showed that phosphorylation at either site is sufficient to protect prostate cancer cells from apoptosis.19, 20, 21 We also showed that BAD promotes prostate tumor growth in mouse models.22 Clinically, while BAD expression was associated with relapse in tamoxifen-treated breast cancer patients,23, 24 phospho-BAD expression was associated with cisplatin resistance and poor overall survival in ovarian cancer.25Our previous findings along with other reports showing the role of BAD in the apoptosis modulation and growth of DC cells19, 22, 26 prompted us to explore the potential role of BAD in the biology of CSCs. We started our investigation by assessing the role of BAD in survival and self-renewal of CSCs. As we observed a significant role for BAD in CSC''s biology, we extended our work to assess the BAD phosphorylation in CSCs of breast cancer patient tumors and for a potential correlation between BAD expression and disease progression in prostate cancer.  相似文献   

11.
This paper presents evidence that a member of the L1 family of ankyrin-binding cell adhesion molecules is a substrate for protein tyrosine kinase(s) and phosphatase(s), identifies the highly conserved FIGQY tyrosine in the cytoplasmic domain as the principal site of phosphorylation, and demonstrates that phosphorylation of the FIGQY tyrosine abolishes ankyrin-binding activity. Neurofascin expressed in neuroblastoma cells is subject to tyrosine phosphorylation after activation of tyrosine kinases by NGF or bFGF or inactivation of tyrosine phosphatases with vanadate or dephostatin. Furthermore, both neurofascin and the related molecule Nr-CAM are tyrosine phosphorylated in a developmentally regulated pattern in rat brain. The FIGQY sequence is present in the cytoplasmic domains of all members of the L1 family of neural cell adhesion molecules. Phosphorylation of the FIGQY tyrosine abolishes ankyrin binding, as determined by coimmunoprecipitation of endogenous ankyrin and in vitro ankyrin-binding assays. Measurements of fluorescence recovery after photobleaching demonstrate that phosphorylation of the FIGQY tyrosine also increases lateral mobility of neurofascin expressed in neuroblastoma cells to the same extent as removal of the cytoplasmic domain. Ankyrin binding, therefore, appears to regulate the dynamic behavior of neurofascin and is the target for regulation by tyrosine phosphorylation in response to external signals. These findings suggest that tyrosine phosphorylation at the FIGQY site represents a highly conserved mechanism, used by the entire class of L1-related cell adhesion molecules, for regulation of ankyrin-dependent connections to the spectrin skeleton.Vertebrate L1, neurofascin, neuroglial cell adhesion molecule (Ng-CAM),1 Ng-CAM–related cell adhesion molecule (Nr-CAM), and Drosophila neuroglian are members of a family of nervous system cell adhesion molecules that possess variable extracellular domains comprised of Ig and fibronectin type III domains and a relatively conserved cytoplasmic domain (Grumet, 1991; Hortsch and Goodman, 1991; Rathgen and Jessel, 1991; Sonderegger and Rathgen, 1992; Hortsch, 1996). Members of this family, including a number of alternatively spliced forms, are abundant in the nervous system during early development as well as in adults. Neurofascin and Nr-CAM, for example, constitute ∼0.5% of the total membrane protein in adult brain (Davis et al., 1993; Davis and Bennett, 1994). Cellular functions attributed to the L1 family include axon fasciculation (Stallcup and Beasley, 1985; Landmesser et al., 1988; Brummendorf and Rathjen, 1993; Bastmeyer et al., 1995; Itoh et al., 1995; Magyar-Lehmann et al., 1995), axonal guidance (van den Pol and Kim, 1993; Liljelund et al., 1994; Brittis and Silver, 1995; Brittis et al., 1995; Lochter et al., 1995; Wong et al., 1996), neurite extension (Chang et al., 1987; Felsenfeld et al., 1994; Hankin and Lagenaur, 1994; Ignelzi et al., 1994; Williams et al., 1994a ,b,c,d; Doherty et al., 1995; Zhao and Siu, 1995), a role in long term potentiation (Luthl et al., 1994), synaptogenesis (Itoh et al., 1995), and myelination (Wood et al., 1990). The potential clinical importance of this group of proteins has been emphasized by the findings that mutations in the L1 gene on the X chromosome are responsible for developmental anomalies including hydrocephalus and mental retardation (Rosenthal et al., 1992; Jouet et al., 1994; Wong et al., 1995).The conserved cytoplasmic domains of L1 family members include a binding site for the membrane skeletal protein ankyrin. This interaction was first described for neurofascin (Davis et. al., 1993) and subsequently has been observed for L1, Nr-CAM (Davis and Bennett, 1994), and Drosophila neuroglian (Dubreuil et al., 1996). The membrane-binding domain of ankyrin contains two distinct sites for neurofascin and has the potential to promote lateral association of neurofascin and presumably other L1 family members (Michaely and Bennett, 1995). Nodes of Ranvier are physiologically relevant axonal sites where ankyrin and L1 family members collaborate, based on findings of colocalization of a specialized isoform of ankyrin with alternatively spliced forms of neurofascin and NrCAM in adults (Davis et al., 1996) as well as in early axonal developmental intermediates (Lambert, S., J. Davis, P. Michael, and V. Bennett. 1995. Mol. Biol. Cell. 6:98a).L1, after homophilic and/or heterophilic binding, participates in signal transduction pathways that ultimately are associated with neurite extension and outgrowth (Ignelzi et al., 1994; Williams et al., 1994a ,b,c,d; Doherty et al., 1995). L1 copurifies with a serine–threonine protein kinase (Sadoul et al., 1989) and is phosphorylated on a serine residue that is not conserved among other family members (Wong et al., 1996). L1 pathway(s) may also involve G proteins, calcium channels, and tyrosine phosphorylation (Williams et al., 1994a ,b,c,d; Doherty et al., 1995). After homophilic interactions, L1 directly activates a tyrosine signaling cascade after a lateral association of its ectodomain with the fibroblast growth factor receptor (Doherty et al., 1995). Antibodies against L1 have also been shown to activate protein tyrosine phosphatase activity in growth cones (Klinz et al., 1995). However, details of the downstream substrates of L1-promoted phosphorylation and dephosphorylation and possible roles of the cytoplasmic domain are not known.Tyrosine phosphorylation is well established to modulate cell–cell and cell–extracellular matrix interactions involving integrins and their associated proteins (Akiyama et al., 1994; Arroyo et al., 1994; Schlaepfer et al., 1994; Law et al., 1996) as well as the cadherins (Balsamo et al., 1996; Krypta et al., 1996; Brady-Kalnay et al., 1995; Shibamoto et al., 1995; Hoschuetzky et al., 1994; Matsuyoshi et al., 1992). For example, the adhesive functions of the calciumdependent cadherin cell adhesion molecule are mediated by a dynamic balance between tyrosine phosphorylation of β-catenin by TrkA and dephosphorylation via the LARtype protein tyrosine phosphatase (Krypta et al., 1996). In this example the regulation of binding among the structural proteins is the result of a coordination between classes of protein kinases and protein phosphatases.This study presents evidence that neurofascin, expressed in a rat neuroblastoma cell line, is a substrate for both tyrosine kinases and protein tyrosine phosphatases at a tyrosine residue conserved among all members of the L1 family. Site-specific tyrosine phosphorylation promoted by both tyrosine kinase activators (NGF and bFGF) and protein tyrosine phosphatase inhibitors (dephostatin and vanadate) is a strong negative regulator of the neurofascin– ankyrin binding interaction and modulates the membrane dynamic behavior of neurofascin. Furthermore, neurofascin and, to a lesser extent Nr-CAM, are also shown here to be tyrosine phosphorylated in developing rat brain, implying a physiological relevance to this phenomenon. These results indicate that neurofascin may be a target for the coordinate control over phosphorylation that is elicited by protein kinases and phosphatases during in vivo tyrosine phosphorylation cascades. The consequent decrease in ankyrin-binding capacity due to phosphorylation of neurofascin could represent a general mechanism among the L1 family members for regulation of membrane–cytoskeletal interactions in both developing and adult nervous systems.  相似文献   

12.
Encapsulated Klebsiella pneumoniae is the predominant causative agent of pyogenic liver abscess, an emerging infectious disease that often complicates metastatic meningitis or endophthalmitis. The capsular polysaccharide on K. pneumoniae surface was determined as the key to virulence. Although the regulation of capsular polysaccharide biosynthesis is largely unclear, it was found that protein-tyrosine kinases and phosphatases are involved. Therefore, the identification and characterization of such kinases, phosphatases, and their substrates would advance our knowledge of the underlying mechanism in capsule formation and could contribute to the development of new therapeutic strategies. Here, we analyzed the phosphoproteome of K. pneumoniae NTUH-K2044 with a shotgun approach and identified 117 unique phosphopeptides along with 93 in vivo phosphorylated sites corresponding to 81 proteins. Interestingly, three of the identified tyrosine phosphorylated proteins, namely protein-tyrosine kinase (Wzc), phosphomannomutase (ManB), and undecaprenyl-phosphate glycosyltransferase (WcaJ), were found to be distributed in the cps locus and thus were speculated to be involved in the converging signal transduction of capsule biosynthesis. Consequently, we decided to focus on the lesser studied ManB and WcaJ for mutation analysis. The capsular polysaccharides of WcaJ mutant (WcaJY5F) were dramatically reduced quantitatively, and the LD50 increased by 200-fold in a mouse peritonitis model compared with the wild-type strain. However, the capsular polysaccharides of ManB mutant (ManBY26F) showed no difference in quantity, and the LD50 increased by merely 6-fold in mice test. Our study provided a clear trend that WcaJ tyrosine phosphorylation can regulate the biosynthesis of capsular polysaccharides and result in the pathogenicity of K. pneumoniae NTUH-K2044.Protein phosphorylation is one of the most biologically relevant and ubiquitous post-translational modifications in both eukaryotic and prokaryotic organisms. It is best known that protein phosphorylation is a reversible enzyme-catalyzed process that is controlled by various kinases and phosphatases. The aberrant functions often result in irregular protein phosphorylation and ultimately lead to serious disease states such as malignant transformation, immune disorders, and pathogenic infections in mammals (1, 2). Recently, accumulating evidences suggest that Ser/Thr/Tyr phosphorylations also contribute to regulate a diverse range of cellular responses and physiological processes in prokaryotes (1). Among them, tyrosine phosphorylation in encapsulated bacteria has been discovered to play key roles in capsular polysaccharide (CPS1; K antigen) biosynthesis, which leads to virulence (3, 4). This thick layer of exopolysaccharide on many pathogenic bacteria can act as a physical boundary to evade phagocytosis and complement-mediated killing and further inhibit complement activation of the host (1, 5, 6).In 1996, Acinetobacter johnsonii protein-tyrosine kinase (Ptk) was first discovered and categorized under the bacterial protein-tyrosine kinase (BY-kinase) family (1, 7, 8). Shortly after, its function in bacterial exopolysaccharide production and transport was characterized (1, 7, 8). From then on, many more bacterial tyrosine kinases such as Wzc of Escherichia coli (1, 9) and EpsB of Pseudomonas solanacearum (10, 11) were found to possess this conserved property; deletion of such tyrosine kinases will result in the loss of exopolysaccharide production (12). Therefore, several experiments were conducted to investigate the role of the downstream substrates of the tyrosine kinases in different strains of bacteria, and some targeted proteins were found to participate in the exopolysaccharide anabolism (13, 14). These findings demonstrated a direct relationship between bacterial tyrosine phosphorylation and exopolysaccharide biosynthesis that was directly reflected in the strain virulence.In the past, the functional roles of the critical components involved in protein phosphorylation were defined by basic biochemical and genetic approaches (1). However, there exists a salient gap between the growing number of identified protein-tyrosine kinases/phosphatases and the relative paucity of protein substrates characterized to date. Genomic sequence analyses and advanced high resolution/high accuracy MS systems with vastly improved phosphopeptide enrichment strategies are among the two key enabling technologies that allow a high efficiency identification of the scarcely detectable site-specific phosphorylations in bacterial systems (15). Mann et al. (16) were the first to initiate a systematic study of the phosphoproteome of B. subtilis in 2007 followed by similar site-specific phosphoproteomics analyses of E. coli (17), Lactococcus lactis (18), and Halobacterium salinarum (19). These pioneering works have since set the foundation in bacterial phosphoproteomics but have not been specifically carried out to address a particular biological issue of causal relevance to virulence or pathogenesis.Klebsiella pneumoniae is a Gram-negative, non-motile, facultative anaerobic, and rod-shaped bacterium. It is commonly found in water and soil (20) as well as on plants (21) and mucosal surfaces of mammals, such as human, horse, and swine (22, 23). It was demonstrated that CPS on the surface of K. pneumoniae is the prime factor of virulence and toxicity in causing pyogenic liver abscess (PLA), a common intra-abdominal infection with a high 10–30% mortality rate worldwide (2429). There are also variations in virulence in regard to different capsular serotypes; K1 and K2 were found to be especially pathogenic in causing PLA in a mouse model (30) compared with other serotypes, which show little or no effect (3134). The K. pneumoniae NTUH-K2044 (K2044) strain, encapsulated with K1 antigen (35), was isolated from clinical K. pneumoniae liver abscess patients. It has become an important emerging pathogen (36) because it usually complicates metastatic septic endophthalmitis and irreversible central nervous system infections independent of host underlying diseases (30, 34). The transmission rate is high (37), and it often rapidly leads to outbreaks of community-acquired infections, such as bacteremia, nosocomial pneumonia, and sepsis, common in immunocompromised individuals (38).In this study, we wanted to prove that the biosynthesis of CPS is mediated through tyrosine phosphorylation of a subset of proteins. An MS-based systematic phosphoproteomics analysis was conducted on K2044 to identify tyrosine phosphorylated proteins that are also associated with CPS biosynthesis. We further validated the relationship between tyrosine phosphorylation on those proteins and virulence of K2044 by site-directed mutagenesis, CPS quantification, serum killing, and mouse lethality assay.  相似文献   

13.
To quantify cell cycle-dependent fluctuations on a proteome-wide scale, we performed integrative analysis of the proteome and phosphoproteome during the four major phases of the cell cycle in Schizosaccharomyces pombe. In highly synchronized cells, we identified 3753 proteins and 3682 phosphorylation events and relatively quantified 65% of the data across all phases. Quantitative changes during the cell cycle were infrequent and weak in the proteome but prominent in the phosphoproteome. Protein phosphorylation peaked in mitosis, where the median phosphorylation site occupancy was 44%, about 2-fold higher than in other phases. We measured copy numbers of 3178 proteins, which together with phosphorylation site stoichiometry enabled us to estimate the absolute amount of protein-bound phosphate, as well as its change across the cell cycle. Our results indicate that 23% of the average intracellular ATP is utilized by protein kinases to phosphorylate their substrates to drive regulatory processes during cell division. Accordingly, we observe that phosphate transporters and phosphate-metabolizing enzymes are phosphorylated and therefore likely to be regulated in mitosis.Cell replication involves a complex series of highly regulated and evolutionary conserved events, called the “cell cycle.” Aberrations in the cell cycle have severe implications and can cause cancerous growth. A detailed understanding of the cell cycle and its regulation may identify additional targets for cancer therapy (13). The cell cycle has been the subject of previous proteomics studies. Olsen et al. (4) measured the dynamics of thousands of proteins and phosphorylation events across cell cycle phases of HeLa cells, providing insights into the underlying regulatory mechanisms and pointing to a general increase in phosphorylation site occupancy during M phase. In a targeted study, Pagliuca et al. (5) investigated interactors of cyclins E1, A2, and B1 in HeLa cells, revealing key mechanistic links between DNA replication and mitosis.Schizosaccharomyces pombe (fission yeast) is a unicellular organism, which can easily be genetically manipulated and carries many cell cycle features similar to metazoan cells. It is an important model organism to study the cell cycle and its checkpoint controls (6). Recent global proteomics studies of yeasts and their cell cycle (713) have mainly focused on Saccharomyces cerevisiae (budding yeast), with only a few studies of fission yeast (14, 15), although the fission yeast cell cycle may be more representative of eukaryotic cell cycles (16). However, attention of the proteomics community toward S. pombe is increasing. Recent proteomics studies covered up to 4087 S. pombe proteins (71% of the predicted proteome) and 1544 phosphoproteins in both asynchronous and synchronized cell cultures (1722); however, a comprehensive analysis of the S. pombe cell cycle is so far missing.Here, we use high resolution mass spectrometry in combination with stable isotope labeling by amino acids in the cell culture (SILAC)1 method, termed super-SILAC (23), and intensity-based absolute quantification (iBAQ) (24) to measure relative and absolute dynamics of the proteome and phosphoproteome during the cell cycle of fission yeast. We estimate copy numbers for 3178 S. pombe proteins, and we combine these data with calculated phosphorylation site stoichiometry to estimate the total amount of protein-bound phosphate and its dynamics across the cell cycle. Providing the global absolute dynamics and stoichiometry of proteins and their modifications will be a valuable resource for classical and systems biologists alike.  相似文献   

14.
Collapsin response mediator protein 2 (CRMP2) is an intracellular protein that mediates signaling of Semaphorin3A (Sema3A), a repulsive axon guidance molecule. Fyn, a Src-type tyrosine kinase, is involved in the Sema3A signaling. However, the relationship between CRMP2 and Fyn in this signaling pathway is still unknown. In our research, we demonstrated that Fyn phosphorylated CRMP2 at Tyr32 residues in HEK293T cells. Immunohistochemical analysis using a phospho-specific antibody at Tyr32 of CRMP showed that Tyr32-phosphorylated CRMP was abundant in the nervous system, including dorsal root ganglion neurons, the molecular and Purkinje cell layer of adult cerebellum, and hippocampal fimbria. Overexpression of a nonphosphorylated mutant (Tyr32 to Phe32) of CRMP2 in dorsal root ganglion neurons interfered with Sema3A-induced growth cone collapse response. These results suggest that Fyn-dependent phosphorylation of CRMP2 at Tyr32 is involved in Sema3A signaling.Collapsin response mediator proteins (CRMPs)4 have been identified as intracellular proteins that mediate Semaphorin3A (Sema3A) signaling in the nervous system (1). CRMP2 is one of the five members of the CRMP family. CRMPs also mediate signal transduction of NT3, Ephrin, and Reelin (24). CRMPs interact with several intracellular molecules, including tubulin, Numb, kinesin1, and Sra1 (58). CRMPs are involved in axon guidance, axonal elongation, cell migration, synapse maturation, and the generation of neuronal polarity (1, 2, 4, 5).CRMP family proteins are known to be the major phosphoproteins in the developing brain (1, 9). CRMP2 is phosphorylated by several Ser/Thr kinases, such as Rho kinase, cyclin-dependent kinase 5 (Cdk5), and glycogen synthase kinase 3β (GSK3β) (2, 1013). The phosphorylation sites of CRMP2 by these kinases are clustered in the C terminus and have already been identified. Rho kinase phosphorylates CRMP2 at Thr555 (10). Cdk5 phosphorylates CRMP2 at Ser522, and this phosphorylation is essential for sequential phosphorylations by GSK3β at Ser518, Thr514, and Thr509 (2, 1113). These phosphorylations disrupt the interaction of CRMP2 with tubulin or Numb (2, 3, 13). The sequential phosphorylation of CRMP2 by Cdk5 and GSK3β is an essential step in Sema3A signaling (11, 13). Furthermore, the neurofibrillary tangles in the brains of people with Alzheimer disease contain hyperphosphorylated CRMP2 at Thr509, Ser518, and Ser522 (14, 15).CRMPs are also substrates of several tyrosine kinases. The phosphorylation of CRMP2 by Fes/Fps and Fer has been shown to be involved in Sema3A signaling (16, 17). Phosphorylation of CRMP2 at Tyr479 by a Src family tyrosine kinase Yes regulates CXCL12-induced T lymphocyte migration (18). We reported previously that Fyn is involved in Sema3A signaling (19). Fyn associates with PlexinA2, one of the components of the Sema3A receptor complex. Fyn also activates Cdk5 through the phosphorylation at Tyr15 of Cdk5 (19). In dorsal root ganglion (DRG) neurons from fyn-deficient mice, Sema3A-induced growth cone collapse response is attenuated compared with control mice (19). Furthermore, we recently found that Fyn phosphorylates CRMP1 and that this phosphorylation is involved in Reelin signaling (4). Although it has been shown that CRMP2 is involved in Sema3A signaling (1, 11, 13), the relationship between Fyn and CRMP2 in Sema3A signaling and the tyrosine phosphorylation site(s) of CRMPs remain unknown.Here, we show that Fyn phosphorylates CRMP2 at Tyr32. Using a phospho-specific antibody against Tyr32, we determined that the residue is phosphorylated in vivo. A nonphosphorylated mutant CRMP2Y32F inhibits Sema3A-induced growth cone collapse. These results indicate that tyrosine phosphorylation by Fyn at Tyr32 is involved in Sema3A signaling.  相似文献   

15.
16.
17.
18.
It has recently become apparent that the Type VI secretion system (T6SS) is a complex macromolecular machine used by many bacterial species to inject effector proteins into eukaryotic or bacterial cells, with significant implications for virulence and interbacterial competition. “Antibacterial” T6SSs, such as the one elaborated by the opportunistic human pathogen, Serratia marcescens, confer on the secreting bacterium the ability to rapidly and efficiently kill rival bacteria. Identification of secreted substrates of the T6SS is critical to understanding its role and ability to kill other cells, but only a limited number of effectors have been reported so far. Here we report the successful use of label-free quantitative mass spectrometry to identify at least eleven substrates of the S. marcescens T6SS, including four novel effector proteins which are distinct from other T6SS-secreted proteins reported to date. These new effectors were confirmed as antibacterial toxins and self-protecting immunity proteins able to neutralize their cognate toxins were identified. The global secretomic study also unexpectedly revealed that protein phosphorylation-based post-translational regulation of the S. marcescens T6SS differs from that of the paradigm, H1-T6SS of Pseudomonas aeruginosa. Combined phosphoproteomic and genetic analyses demonstrated that conserved PpkA-dependent threonine phosphorylation of the T6SS structural component Fha is required for T6SS activation in S. marcescens and that the phosphatase PppA can reverse this modification. However, the signal and mechanism of PpkA activation is distinct from that observed previously and does not appear to require cell–cell contact. Hence this study has not only demonstrated that new and species-specific portfolios of antibacterial effectors are secreted by the T6SS, but also shown for the first time that PpkA-dependent post-translational regulation of the T6SS is tailored to fit the needs of different bacterial species.Gram-negative bacteria have evolved several specialized protein secretion systems to secrete a wide variety of substrate proteins into the extracellular milieu or to inject them into other, often eukaryotic, cells (1). Secreted proteins and their associated secretion systems are very important in bacterial virulence and interactions with other organisms (2). One of the most recent discoveries in this field is the Type VI secretion system (T6SS),1 which occurs widely across bacterial species (3, 4) and can target proteins to both bacterial and eukaryotic cells (5). The significance of the T6SS is becoming increasingly apparent. It has been implicated in virulence, commensalism, and symbiosis with eukaryotes (5, 6). Additionally, in many bacteria, the T6SS is now implicated in antibacterial activity. T6SS-mediated antibacterial killing appears to be important for competition between bacterial species, for example within the resident microflora of a eukaryotic host (5, 7).Secretion by the T6SS relies on 13 conserved core components which are predicted to form a large machinery associated with the cell envelope, including membrane-bound and bacteriophage tail-like subassemblies (8, 9). The membrane bound subassembly consists of inner membrane proteins (TssLM) and an outer membrane lipoprotein (TssJ) and is anchored to the cell wall. The phage tail-like assembly consists of several proteins that show structural homology with T4 phage tail proteins or are organized in similar structures (10). Hcp (TssD) proteins form hexameric rings and are thought to stack into tube-like structures (11, 12). This Hcp tube is believed to be capped by a trimer of VgrG (TssI) proteins, which share structural homology with the needle of the T4 phage tail (10, 13). In addition, VipA (TssB) and VipB (TssC) form a large tubular structure highly reminiscent of the T4 phage tail sheath (14, 15). Such similarities have led to the idea that the T6SS resembles an inverted contractile bacteriophage infection machinery and injects substrates via an Hcp/VgrG needle into other cells. Recent models propose that the VipA/B sheath surrounds the Hcp/VgrG needle and contraction of the VipA/B tube pushes the Hcp/VgrG needle out of the cell (1618). It has been postulated that this mechanism can be triggered by close contact with other neighboring cells (1921).Assembly, localization, and remodelling of VipA/B tubules in vivo depend on the AAA+ ATPase ClpV (TssH), another essential core component of the T6SS (14, 16, 17). ClpV also interacts with the accessory component Fha (TagH) (22, 23), which is found in a subset of T6SSs (4). The Fha protein has an N-terminal domain with a forkhead associated motif, which is predicted to bind phospho-threonine peptides (24). In Pseudomonas aeruginosa, Fha1 is phosphorylated by the Thr/Ser kinase PpkA (TagE) and dephosphorylated by the phosphatase PppA (TagG), and the phosphorylation state of Fha1 regulates the activity of the T6SS (22, 23). Phosphorylation of Fha in P. aeruginosa is also controlled by additional components, which act upstream of PpkA and form a regulatory cascade for T6SS activation (22, 25). Although homologs of PpkA and PppA have been identified in the T6SS gene clusters of certain other bacteria (3), the regulation of the T6SS by post-translational protein phosphorylation has not yet been experimentally investigated outside of Pseudomonas.To understand how the T6SS affects eukaryotic and bacterial cells, it is critical to identify substrate proteins secreted by the T6SS. The VgrG and Hcp proteins were the first identified T6SS substrates and appear to be generally secreted to the external milieu by all T6SSs (26). However, as mentioned above, Hcp and VgrG are core components of the T6SS machinery and therefore represent extracellular components of the secretion apparatus rather than genuine secreted effector proteins. Nonetheless, a limited number of VgrG homologs with extra functional effector domains at the C terminus have been identified or predicted, which account for some of the T6SS dependent effects seen against bacteria and eukaryotes. For example, the C-terminal domain of VgrG-1 from Vibrio cholerae shows actin crosslinking activity in eukaryotic cells (13, 27) and the C-terminal domain of V. cholerae VgrG-3 has bacterial cell wall hydrolase activity (28, 29).Recently, following much effort in the field, a small number of proteins secreted by the T6SS, but not structural components, have been experimentally identified. These proteins are regarded as true secreted substrates of the T6SS, with effector functions in target cells (2935). For example, antibacterial T6SS-secreted effector proteins with peptidoglycan amidase (cell wall hydrolysis) function, the Type VI amidase effector (Tae) proteins, have been identified in Burkholderia thailandensis (32), P. aeruginosa (31), and Serratia marcescens (30). These Tae proteins play a role in T6SS-mediated antibacterial killing activity and genes encoding four families of Tae protein have been widely identified in other bacteria with T6SSs (32). T6SS-secreted effector proteins which are not peptidoglycan hydrolases have also been reported, including Tse2 secreted by P. aeruginosa, which acts in the bacterial cytoplasm (31), and the VasX and TseL proteins secreted by the V. cholerae T6SS, which are suggested to target membrane lipids (29, 34, 35). In the case of antibacterial T6SSs, the secreting bacterial cells are protected from their own T6SS effector proteins by specific immunity proteins (2932, 35). However, given the large number of T6SSs in different bacterial species and their apparent ability to secrete multiple substrates, experimentally identified T6-secreted effector proteins still remain surprisingly scarce.Here we report the identification of multiple T6SS-secreted effector proteins in S. marcescens. S. marcescens is an opportunistic pathogen, for example causing ocular infections, nosocomial septicemia and pneumonia (36). Previously, we have identified a T6SS in S. marcescens Db10, which targets and efficiently kills other bacterial cells and plays a role in antibacterial competition (37). We have recently demonstrated that this T6SS secretes two antibacterial effectors, the Tae4 homologs Ssp1 and Ssp2, with cognate immunity proteins Rap1a and Rap2a (30).In this work, we report the analysis of the T6SS-dependent secretome of S. marcescens by label-free quantitation (LFQ) mass spectrometry and describe the identification and characterization of four novel T6SS-secreted effector proteins. These were confirmed as antibacterial toxins and specific immunity proteins were identified. Additionally, this global secretomic analysis, in combination with genetic and phosphoproteomic analyses, demonstrated that a post-translational phosphorylation system influences the ability of the S. marcescens T6SS to secrete effector proteins. Although this system uses homologs of the P. aeruginosa PpkA, PppA and Fha components, the circumstances and impact of Fha phosphorylation were shown to vary between organisms.  相似文献   

19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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