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

2.
The impact of Nα-terminal acetylation on protein stability and protein function in general recently acquired renewed and increasing attention. Although the substrate specificity profile of the conserved enzymes responsible for Nα-terminal acetylation in yeast has been well documented, the lack of higher eukaryotic models has hampered the specificity profile determination of Nα-acetyltransferases (NATs) of higher eukaryotes. The fact that several types of protein N termini are acetylated by so far unknown NATs stresses the importance of developing tools for analyzing NAT specificities. Here, we report on a method that implies the use of natural, proteome-derived modified peptide libraries, which, when used in combination with two strong cation exchange separation steps, allows for the delineation of the in vitro specificity profiles of NATs. The human NatA complex, composed of the auxiliary hNaa15p (NATH/hNat1) subunit and the catalytic hNaa10p (hArd1) and hNaa50p (hNat5) subunits, cotranslationally acetylates protein N termini initiating with Ser, Ala, Thr, Val, and Gly following the removal of the initial Met. In our studies, purified hNaa50p preferred Met-Xaa starting N termini (Xaa mainly being a hydrophobic amino acid) in agreement with previous data. Surprisingly, purified hNaa10p preferred acidic N termini, representing a group of in vivo acetylated proteins for which there are currently no NAT(s) identified. The most prominent representatives of the group of acidic N termini are γ- and β-actin. Indeed, by using an independent quantitative assay, hNaa10p strongly acetylated peptides representing the N termini of both γ- and β-actin, and only to a lesser extent, its previously characterized substrate motifs. The immunoprecipitated NatA complex also acetylated the actin N termini efficiently, though displaying a strong shift in specificity toward its known Ser-starting type of substrates. Thus, complex formation of NatA might alter the substrate specificity profile as compared with its isolated catalytic subunits, and, furthermore, NatA or hNaa10p may function as a post-translational actin Nα-acetyltransferase.The multisubunit and ribosome-associated protein Nα-acetyltransferases (NATs)1 are omnipresent enzyme complexes that catalyze the transfer of the acetyl moiety from acetyl-CoA to the primary α-amines of N termini of nascent proteins (13). As up to 50 to 60% of yeast proteins and 80 to 90% of human proteins are modified in this manner, Nα-acetylation is a widespread protein modification in eukaryotes (47), and the pattern of modification has remained largely conserved throughout evolution (4, 8). NATs belong to a subfamily of the Gcn5-related N-acetyltransferase superfamily of N-acetyltransferases, additionally encompassing the well-studied histone acetyltransferases that are implicated in epigenetic imprinting.In yeast and humans, three main NAT complexes, NatA, NatB, and NatC were found to be responsible for the majority of Nα-terminal acetylations (1). The NatA complex, responsible for cotranslational Nα-terminal acetylation of proteins with Ser, Ala, Thr, Gly, and Val N termini, is composed of two main subunits, the catalytic subunit Naa10p (previously known as Ard1p) and the auxiliary subunit Naa15p (previously known as Nat1p/NATH) (911). Furthermore, a third catalytic subunit Naa50p (previously known as Nat5)—an acetyltransferase shown to function in chromosome cohesion and segregation (1214)—was found to physically interact with the NatA complex of yeast (2), fruit fly (12), and human (15). Recently, human Naa50p (hNaa50p) was reported to display lysine or Nε-acetyltransferase as well as NAT activity (16), the latter was defined as NatE activity (16). Interestingly, the chaperone-like, Huntingtin interacting protein HYPK, identified as a novel stable interactor of human NatA, was functionally implicated in the N-terminal acetylation of an in vivo NatA substrate, demonstrating that NAT complex formation and composition may have an overall influence on the observed (degree of) Nα-acetylation (17). Further, subunits of the human NatA complex have been coupled to cancer-related processes and differentiation, with altered subunit expression reported in papillary thyroid carcinoma, neuroblastoma, and retinoic acid induced differentiation. Furthermore, the NatA catalytic subunit was found to be implicated in processes such as hypoxia-response and the β-catenin pathway (18, 19). Of note is that in line with the differential localization patterns of the individual NatA subunits (9, 13, 20, 21), other data indicate that these subunits might well exert NatA-independent enzymatic functions (13, 22, 23). Given that a significant fraction of hNaa10p and hNaa15p are nonribosomal (9), and given the multitude of postulated post-translational in vivo N-acetylation events recently reported (2426), these observations argue in favor of the existence of NAT complexes and/or catalytic NAT-subunits acting post-translationally.Similar to NatA, the NatB and NatC complexes, composed of the catalytic subunit Naa20p or Naa30p and the auxiliary subunits Naa25p or Naa35p and Naa38p respectively, are conserved from yeast to higher eukaryotes concerning their subunit composition as well as their substrate specificity. Both these complexes display activity toward methionine-starting N termini, with NatB preferring acidic residues as well as Asn and Gln at P2′-sites2, whereas NatC prefers hydrophobic amino acid residues at substrate P2′-sites (1, 27, 28).Nα-acetylation affects various protein functions such as localization, activity, association, and stability (29, 30). Only recently a more generalized function of protein Nα-acetylation in generating so-called N-terminal degrons marking proteins for removal was put forward (31). The lack of mouse models in addition to the fact that (combined) knockdown of individual components of Nα-acetyltransferases only marginally affect the overall Nα-acetylation status (4) have so far hampered the molecular characterization of the substrate specificity profile of (yet uncharacterized) NATs. To date, all eukaryote Nα-acetylation events are assumed to be catalyzed by the five known NATs (32). However, an additional level of complexity is imposed by the fact that in contrast to yeast, higher eukaryotes express multiple splice variants of various NAT subunits as well as paralogs thereof (33, 34), further implicating that a specific NAT''s substrate specificity might be altered in this way, in addition to the possible existence of substrate redundancy. Moreover, regulation of substrate specificity and stability of NAT activity can be imposed by differential complex formation and post-translational modifications including phosphorylation, auto-acetylation, and specific proteolytic cleavage of the catalytic subunits (9, 16, 17). As such, a detailed understanding of the substrate specificity of NATs, and the regulation thereof, could help unravel the physiological substrate repertoires as well as the associated physiological roles of NATs in the normal and the disease state.The specificity of Nα-acetyltransferases and their endogenous substrates were originally studied by two-dimensional-PAGE: Nα-acetylation neutralizes the N-terminal positive charge, resulting in an altered electrophoretic protein migration during isoelectric focusing (3538). Recently, this altered biophysical property was also exploited to enrich for protein N-termini using low pH strong cation exchange (SCX) chromatography (24, 39). As an example, SCX prefractionation combined with N-terminal combined fractional diagonal chromatography, a targeted proteomics technology negatively selecting for protein N-terminal peptides, stable isotope labeling of amino acids in cell culture, and amino-directed modifiers (40), was used to study the in vivo substrate repertoires of human as well as yeast NatA (4).Nevertheless, the various methods reported today to study in detail Nα-terminal acetylation and thus the specificities of different NATs make use of a limited and therefore somewhat biased set of synthesized peptide substrates and comprise the rather laborious detection of radioactive acetylated products as well as enzyme-coupled methods quantifying acetyl-CoA conversion. Because (proteome-derived) peptide libraries have been used extensively to study epitope mapping (41), protein-protein interactions (42), protein modifications such as phosphorylation (43), and proteolysis (44, 45), as well as for determining the substrate specificity of the Nα-deblocking peptide deformylase (46), we reckoned that the development of an oligopeptide-based acetylation assay should allow for more comprehensive screening of NAT-like activities. We here report on the development of a peptide-based method to systematically screen for the in vitro sequence specificity profile of individual NATs as well as endogenous NAT complexes. In summary, SCX enriched, Nα-free peptide libraries, derived from natural proteomes build up the peptide substrate pool. And, upon incubation, NAT Nα-acetylated peptides are enriched by a second SCX fractionation step, resulting in a positive selection of NAT-specific peptide substrates. By use of this proteome-derived peptide library approach, we here delineated (differences in) the specificity profiles of hNaa50p and hNaa10p as isolated hNatA components, as well as of assayed their combined activity when in their native hNatA complex.  相似文献   

3.
The human NatA protein Nα-terminal-acetyltransferase complex is responsible for cotranslational N-terminal acetylation of proteins with Ser, Ala, Thr, Gly, and Val N termini. The NatA complex is composed of the catalytic subunit hNaa10p (hArd1) and the auxiliary subunit hNaa15p (hNat1/NATH). Using immunoprecipitation coupled with mass spectrometry, we identified endogenous HYPK, a Huntingtin (Htt)-interacting protein, as a novel stable interactor of NatA. HYPK has chaperone-like properties preventing Htt aggregation. HYPK, hNaa10p, and hNaa15p were associated with polysome fractions, indicating a function of HYPK associated with the NatA complex during protein translation. Knockdown of both hNAA10 and hNAA15 decreased HYPK protein levels, possibly indicating that NatA is required for the stability of HYPK. The biological importance of HYPK was evident from HYPK-knockdown HeLa cells displaying apoptosis and cell cycle arrest in the G0/G1 phase. Knockdown of HYPK or hNAA10 resulted in increased aggregation of an Htt-enhanced green fluorescent protein (Htt-EGFP) fusion with expanded polyglutamine stretches, suggesting that both HYPK and NatA prevent Htt aggregation. Furthermore, we demonstrated that HYPK is required for N-terminal acetylation of the known in vivo NatA substrate protein PCNP. Taken together, the data indicate that the physical interaction between HYPK and NatA seems to be of functional importance both for Htt aggregation and for N-terminal acetylation.Nα-terminal acetylation is among the most common protein modifications in eukaryotes, occurring on ∼50% of Saccharomyces cerevisiae proteins and ∼80% of human proteins (12). In yeast, four types of Nα-terminal acetyltransferases (NATs) have been defined (NatA-NatD), while a fifth type, NatE, has been hypothesized (21, 32-34, 38). For humans, NatA, NatB, NatC, and NatE were recently presented (2, 4, 18, 39, 40). A revised NAT-subunit nomenclature was recently introduced in order to have identical names for orthologous subunits from different species, and each gene was denoted NAA (Nα-acetyltransferase) followed by a number depending on Nat type and the type of subunit (catalytic/auxiliary) (32). The major human NAT complex, hNatA, is composed of the catalytic subunit hNaa10p (previously named hArd1) and the auxiliary subunit hNaa15p (hNat1/NATH) (4). Human NatA is evolutionarily conserved from the yeast complex in terms of subunit composition and substrate specificity (12, 26, 28). However, in contrast to yeast cells, human cells potentially contain several distinct NatA complexes due to the presence of two genes for each of the two NatA subunits, NAA10 and NAA15 (6, 8). Protein N-terminal acetylation occurs on the ribosome when the nascent polypeptide emerges (21, 29, 30, 41, 42). Proteins with Ser, Thr, Gly, Ala, Val, or Cys N termini are potential substrates of NatA (12), while NatB and NatC potentially acetylate specific classes of substrates that still carry the initiator Met (34). The biological importance of the human NatA complex was evident from knockdown experiments where induction of apoptosis and growth arrest of cells in the G1/G0 phase were the resulting phenotypes (9, 11, 20, 25). The phenotypes induced by hNatA depletion most likely reflect the fact that one or more specific substrate proteins lack proper Nα acetylation, in view of the fact that a large quantitative proteomic analysis of the acetylation status of protein N termini in hNaa15p-hNaa10p knockdown cells revealed a decrease in the level of Nα acetylation of some partially acetylated substrates compared to that in control cells (12).To further characterize the human NatA complex, we looked for the presence of stable interaction partners of hNaa15p and hNaa10p. Here we present data identifying the Huntingtin (Htt) yeast two-hybrid protein K (HYPK) as a novel factor involved in cotranslational NatA acetylation. HYPK, originally identified in a yeast two-hybrid screen during a search for potential interaction partners for the Huntingtin protein (19), was recently found to reduce Htt polyglutamine (polyQ) aggregation upon overexpression (36). However, the role of the endogenous HYPK protein has yet to be revealed. We demonstrate that endogenous HYPK (i) stably interacts with the hNaa10p-hNaa15p NatA N-terminal-acetyltransferase complex and with ribosomes, (ii) is required for normal N-terminal acetylation of a NatA substrate, (iii) is important for cell survival independent of Htt polyQ, and (iv) is important for the prevention of Htt polyQ aggregation. Furthermore, NatA is essential for the proper expression of HYPK protein and modulates Htt polyQ aggregation.  相似文献   

4.
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   

5.
6.
In Alzheimer disease (AD) and frontotemporal dementia the microtubule-associated protein Tau becomes progressively hyperphosphorylated, eventually forming aggregates. However, how Tau dysfunction is associated with functional impairment is only partly understood, especially at early stages when Tau is mislocalized but has not yet formed aggregates. Impaired axonal transport has been proposed as a potential pathomechanism, based on cellular Tau models and Tau transgenic mice. We recently reported K369I mutant Tau transgenic K3 mice with axonal transport defects that suggested a cargo-selective impairment of kinesin-driven anterograde transport by Tau. Here, we show that kinesin motor complex formation is disturbed in the K3 mice. We show that under pathological conditions hyperphosphorylated Tau interacts with c-Jun N-terminal kinase- interacting protein 1 (JIP1), which is associated with the kinesin motor protein complex. As a result, transport of JIP1 into the axon is impaired, causing JIP1 to accumulate in the cell body. Because we found trapping of JIP1 and a pathological Tau/JIP1 interaction also in AD brain, this may have pathomechanistic implications in diseases with a Tau pathology. This is supported by JIP1 sequestration in the cell body of Tau-transfected primary neuronal cultures. The pathological Tau/JIP1 interaction requires phosphorylation of Tau, and Tau competes with the physiological binding of JIP1 to kinesin light chain. Because JIP1 is involved in regulating cargo binding to kinesin motors, our findings may, at least in part, explain how hyperphosphorylated Tau mediates impaired axonal transport in AD and frontotemporal dementia.The microtubule-associated protein Tau is predominantly found in the axonal compartment of neurons, where it binds to microtubules (1). In human brain, six isoforms of Tau are expressed, due to alternative splicing of exons 2, 3 and 10 (2). Tau consists of an amino-terminal projection domain followed by 3 or 4 microtubule binding repeats (3R or 4R), due to splicing of exon 10, and a carboxyl-terminal tail region. In the AD3 and FTD brain, Tau forms filamentous inclusions (3). They are found in nerve cell bodies and apical dendrites as neurofibrillary tangles (NFTs), in distal dendrites as neuropil threads, and in the abnormal neurites that are associated with some amyloid plaques (neuritic plaques) (3). Hyperphosphorylation of Tau is thought to be an initiating step (4), as it detaches Tau from microtubules and makes it prone to form aggregates (1, 5). Whereas in AD no mutations have been identified in the MAPT gene encoding Tau, so far 42 intronic and exonic mutations have been found in familial forms of FTD (6). Their identification assisted in the generation of transgenic mouse models that reproduce NFT formation and memory impairment (7).The models were also instrumental in testing hypotheses that had been brought forward to link Tau pathology to functional impairment (810). In particular, defects in axonal transport have been implicated in neurodegenerative disorders (11, 12). Tau binding to microtubules affects axonal transport (13), and in cell culture overexpression of Tau was shown to lead to impaired transport of mitochondria and vesicles (14, 15). Axonal transport defects have also been reproduced in wild-type Tau transgenic mice (16) and in K369I mutant Tau K3 mice (17), whereas Tau expression failed to inhibit axonal transport in other systems (18, 19). This apparent discrepancy may depend on the type of cargos analyzed and, specifically, the experimental paradigm, e.g. using phosphorylated (16, 17, 20) versus non-phosphorylated Tau (18).To dissect Tau-mediated axonal transport defects at a molecular level, we used K3 mice that overexpress human Tau carrying the pathogenic FTD K369I mutation (17). We observed a pronounced hyperphosphorylation of transgenic Tau in many brain areas. Clinically, the mice present with an early onset motor phenotype that is, at least in part, caused by impairment of axonal transport in neurons of the substantia nigra. Interestingly, only selected aspects of anterograde axonal transport were impaired, in particular those of kinesin-I motor complex-driven vesicles and mitochondria. Our data suggest a selective impairment of axonal transport rather than a generalized, non-selective blockage of microtubules that has been established in cell culture systems, which fail to phosphorylate Tau at the high levels that are found in vivo even under physiological conditions. More importantly, in AD and FTD Tau is even more phosphorylated, i.e. hyperphosphorylated at physiological sites and de novo at pathological sites, preventing it from binding to microtubules (1).Based on our findings of an impaired kinesin-I-driven axonal transport in the K3 mice, we speculated that hyperphosphorylated Tau may impair anterograde transport by interfering directly with components of the kinesin-I motor complex rather than disrupting the binding of the kinesin heavy chain (see below) to microtubules. Axonal transport along microtubules is mediated by members of the kinesin superfamily (KIF) of motor proteins (2123). The KIFs typically consist of an ATPase domain that interacts with microtubules and drives movement and a domain that links to cargos, either directly or indirectly, as in the case of KIF5, by assembling with the kinesin light chain (KLC) to form the kinesin-I (KIF5/KLC) motor complex (24). In addition, increasing evidence suggests that scaffolding proteins mediate and regulate the binding of cargos to KIFs (21, 2527). These include the scaffold protein JNK-interacting protein (JIP) that is involved in the linkage of cargos to the kinesin-I motor complex via KLC (25, 2833).Here, by using the K3 mouse model, we identified a novel interaction of Tau and JIP in neurons that causes a trapping of JNK interacting protein 1 (JIP1) in the cell body of K3 mice, cell culture systems, and human AD brain. We found that the pathological interaction of hyperphosphorylated Tau and JIP1 competes with the physiological binding of JIP1 to KLC.  相似文献   

7.
8.
9.
Isogenic, E3-deleted adenovirus vectors defective in E1, E1 and E2A, or E1 and E4 were generated in complementation cell lines expressing E1, E1 and E2A, or E1 and E4 and characterized in vitro and in vivo. In the absence of complementation, deletion of both E1 and E2A completely abolished expression of early and late viral genes, while deletion of E1 and E4 impaired expression of viral genes, although at a lower level than the E1/E2A deletion. The in vivo persistence of these three types of vectors was monitored in selected strains of mice with viral genomes devoid of transgenes to exclude any interference by immunogenic transgene-encoded products. Our studies showed no significant differences among the vectors in the short-term maintenance and long-term (4-month) persistence of viral DNA in liver and lung cells of immunocompetent and immunodeficient mice. Furthermore, all vectors induced similar antibody responses and comparable levels of adenovirus-specific cytotoxic T lymphocytes. These results suggest that in the absence of transgenes, the progressive deletion of the adenovirus genome does not extend the in vivo persistence of the transduced cells and does not reduce the antivirus immune response. In addition, our data confirm that, in the absence of transgene expression, mouse cellular immunity to viral antigens plays a minor role in the progressive elimination of the virus genome.Replication-deficient human adenoviruses (Ad) have been widely investigated as ex vivo and in vivo gene delivery systems for human gene therapy. The ability of these vectors to mediate the efficient expression of candidate therapeutic or vaccine genes in a variety of cell types, including postmitotic cells, is considered an advantage over other gene transfer vectors (3, 28, 49). However, the successful application of currently available E1-defective Ad vectors in human gene therapy has been hampered by the fact that transgene expression is only transient in vivo (2, 15, 16, 33, 36, 46). This short-lived in vivo expression of the transgene has been explained, at least in part, by the induction in vivo of cytotoxic immune responses to cells infected with the Ad vector. Studies with rodent systems have suggested that cytotoxic T lymphocytes (CTLs) directed against virus antigens synthesized de novo in the transduced tissues play a major role in eliminating cells containing the E1-deleted viral genome (5658, 61). Consistent with the concept of cellular antiviral immunity, expression of transgenes is significantly extended in experimental rodent systems that are deficient in various components of the cellular immune system or that have been rendered immunocompromised by administration of pharmacological agents (2, 33, 37, 48, 60, 64).Based on the assumption that further reduction of viral antigen expression may lower the immune response and thus extend persistence of transgene expression, previous studies have investigated the consequences of deleting both E1 and an additional viral regulatory region, such as E2A or E4. The E2A region encodes a DNA binding protein (DBP) with specific affinity for single-stranded Ad DNA. The DNA binding function is essential for the initiation and elongation of viral DNA synthesis during the early phase of Ad infection. During the late phase of infection, DBP plays a central role in the activation of the major late promoter (MLP) (for a recent review, see reference 44). The E4 region, located at the right end of the viral genome, encodes several regulatory proteins with pleiotropic functions which are involved in the accumulation, splicing, and transport of early and late viral mRNAs, in DNA replication, and in virus particle assembly (reviewed in reference 44). The simultaneous deletion of E1 and E2A or of E1 and E4 should therefore further reduce the replication of the virus genome and the expression of early and late viral genes. Such multidefective vectors have been generated and tested in vitro and in vivo (9, 12, 17, 1921, 23, 24, 26, 34, 40, 52, 53, 59, 62, 63). Recombinant vectors with E1 deleted and carrying an E2A temperature-sensitive mutation (E2Ats) have been shown in vitro to express much smaller amounts of virus proteins, leading to extended transgene expression in cotton rats and mice (19, 20, 24, 59). To eliminate the risks of reversion of the E2Ats point mutation to a wild-type phenotype, improved vectors with both E1 and E2A deleted were subsequently generated in complementation cell lines coexpressing E1 and E2A genes (26, 40, 63). In vitro analysis of human cells infected by these viruses demonstrated that the double deletion completely abolished viral DNA replication and late protein synthesis (26). Similarly, E1/E4-deleted vectors have been generated in various in vitro complementation systems and tested in vitro and in vivo (9, 17, 23, 45, 52, 53, 62). These studies showed that deletion of both E1 and E4 did indeed reduce significantly the expression of early and late virus proteins (17, 23), leading to a decreased anti-Ad host immune response (23), reduced hepatotoxicity (17, 23, 52), and improved in vivo persistence of the transduced liver cells (17, 23, 52).Interpretation of these results is difficult, however, since all tested E1- and E1/E4-deleted vectors encoded the bacterial β-galactosidase (βgal) marker, whose strong immunogenicity is known to influence the in vivo persistence of Ad-transduced cells (32, 37). Moreover, the results described above are not consistent with the conclusions from other studies showing, in various immunocompetent mouse models, that cellular immunity to Ad antigens has no detectable impact on the persistence of the transduced cells (37, 40, 50, 51). Furthermore, in contrast to results of earlier studies (19, 20, 59), Fang et al. (21) demonstrated that injection of E1-deleted/E2Ats vectors into immunocompetent mice and hemophilia B dogs did not lead to an improvement of the persistence of transgene expression compared to that with isogenic E1-deleted vectors. Similarly, Morral et al. (40) did not observe any difference in persistence of transgene expression in mice injected with either vectors deleted in E1 only or vectors deleted in both E1 and E2A. Finally, the demonstration that some E4-encoded products can modulate transgene expression (1, 17, 36a) makes the evaluation of E1- and E1/E4-deleted vectors even more complex when persistence of transgene expression is used for direct comparison of the in vivo persistence of cells transduced by the two types of vectors.The precise influence of the host immune response to viral antigens on the in vivo persistence of the transduced cells, and hence the impact of further deletions in the virus genome, therefore still remains unclear. To investigate these questions, we generated a set of isogenic vectors with single deletions (AdE1°) and double deletions (AdE1°E2A° and AdE1°E4°) and their corresponding complementation cell lines and compared the biologies and immunogenicities of these vectors in vitro and in vivo. To eliminate any possible influence of transgene-encoded products on the interpretation of the in vivo results, we used E1-, E1/E2A-, and E1/E4-deleted vectors with no transgenes.  相似文献   

10.
11.
12.
13.
14.
Mathematical tools developed in the context of Shannon information theory were used to analyze the meaning of the BLOSUM score, which was split into three components termed as the BLOSUM spectrum (or BLOSpectrum). These relate respectively to the sequence convergence (the stochastic similarity of the two protein sequences), to the background frequency divergence (typicality of the amino acid probability distribution in each sequence), and to the target frequency divergence (compliance of the amino acid variations between the two sequences to the protein model implicit in the BLOCKS database). This treatment sharpens the protein sequence comparison, providing a rationale for the biological significance of the obtained score, and helps to identify weakly related sequences. Moreover, the BLOSpectrum can guide the choice of the most appropriate scoring matrix, tailoring it to the evolutionary divergence associated with the two sequences, or indicate if a compositionally adjusted matrix could perform better.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29]  相似文献   

15.
The U16 small nucleolar RNA (snoRNA) is encoded by the third intron of the L1 (L4, according to the novel nomenclature) ribosomal protein gene of Xenopus laevis and originates from processing of the pre-mRNA in which it resides. The U16 snoRNA belongs to the box C/D snoRNA family, whose members are known to assemble in ribonucleoprotein particles (snoRNPs) containing the protein fibrillarin. We have utilized U16 snoRNA in order to characterize the factors that interact with the conserved elements common to the other members of the box C/D class. In this study, we have analyzed the in vivo assembly of U16 snoRNP particles in X. laevis oocytes and identified the proteins which interact with the RNA by label transfer after UV cross-linking. This analysis revealed two proteins, of 40- and 68-kDa apparent molecular size, which require intact boxes C and D together with the conserved 5′,3′-terminal stem for binding. Immunoprecipitation experiments showed that the p40 protein corresponds to fibrillarin, indicating that this protein is intimately associated with the RNA. We propose that fibrillarin and p68 represent the RNA-binding factors common to box C/D snoRNPs and that both proteins are essential for the assembly of snoRNP particles and the stabilization of the snoRNA.One of the most interesting recent findings related to ribosome biogenesis has been the identification of a large number of small RNAs localized in the nucleolus (snoRNAs). So far, more than 60 snoRNAs have been identified in vertebrates (17), and more than 30 have been identified in yeast (2). The total number of snoRNAs is not known, but it is likely to be close to 200 (33, 38). These snoRNAs, with the exception of the mitochondrial RNA processing (MRP) species (38), can be grouped into two major families on the basis of conserved structural and sequence elements. The first group includes molecules referred to as box C/D snoRNAs, whereas the second one comprises the species belonging to the box H/ACA family (2, 15).The two families differ in many aspects. The box C/D snoRNAs are functionally heterogeneous. Most of them function as antisense RNAs in site-specific ribose methylation of the pre-rRNA (1, 10, 17, 26); a minority have been shown to play a direct role in pre-rRNA processing in both yeast and metazoan cells (11, 21). The box C/D snoRNAs play their role by means of unusually long (up to 21 contiguous nucleotides) regions of complementarity to highly conserved sequences of 28S and 18S rRNAs (1). In contrast, several members of the H/ACA RNA family have been shown to direct site-specific isomerization of uridines into pseudouridines and to display shorter regions of complementarity to rRNA (14, 24). Mutational analysis suggests that H/ACA snoRNAs can also play a role as antisense RNAs by base pairing with complementary regions on rRNA (15, 24).Another difference between the two families can be seen by comparison of secondary structures. A Y-shaped motif, where a 5′,3′-terminal stem adjoins the C and D conserved elements, has been proposed for many box C/D snoRNAs (16, 26, 40, 42), whereas box H/ACA snoRNAs have been proposed to fold into two conserved hairpin structures connected by a single-stranded hinge region, followed by a short 3′ tail (15).Despite these differences, analogies have been found in the roles played by the conserved box elements. Mutational analysis and competition experiments indicated that C/D and H/ACA boxes are required both for processing and stable accumulation of the mature snoRNA, suggesting that they represent binding sites for specific trans-acting factors (2, 3, 8, 15, 16, 28, 36, 41).All snoRNAs are associated with proteins to form specific ribonucleoparticles (snoRNPs). The study of these particles began only recently, and so far, very few aspects of their structure and biosynthesis have been clarified. The only detailed analysis performed was on the mammalian U3 (19) and the yeast snR30 (20) snoRNPs. Of the identified components, a few appear to be more general factors: fibrillarin, which was shown to be associated with C/D snoRNPs (3, 4, 8, 13, 28, 31, 39), and the nucleolar protein GAR1, which was found associated with H/ACA snoRNAs in yeast (20). Just as the study of small nuclear RNP (snRNP) particles was crucial to the understanding of the splicing process, a detailed structural and functional analysis of snoRNP particles will be essential to elucidate the complex process of ribosome biosynthesis.In this study, we have analyzed the snoRNP assembly of wild-type and mutant U16 snoRNAs by following the kinetics of complex formation in the in vivo system of the Xenopus laevis oocyte. By a UV cross-linking technique, we have identified two proteins, of 40- and 68-kDa apparent molecular mass, which require intact boxes C and D together with the terminal stem for their binding. The 40-kDa species is specifically recognized by fibrillarin antibodies, indicating that this protein is intimately associated with the RNA.  相似文献   

16.
17.
18.
The general stress response of Bacillus subtilis can be activated by stimuli such as the addition of salt or ethanol and with blue light. In the latter response, YtvA activates σB through a cascade of Rsb proteins, organized in stressosomes. YtvA is composed of an N-terminal LOV (light, oxygen, and voltage) domain and a C-terminal STAS (sulfate transporter and anti-sigma factor) domain and shows light-modulated GTP binding in vitro. Here, we examine the mechanism of YtvA-mediated activation of σB in vivo with site-directed mutagenesis. Constitutive off and constitutive on mutations have been identified. Disruption of GTP binding in the STAS domain eliminates light activation of σB. In contrast, modification of sites relevant for phosphorylation of STAS domains does not affect the stress response significantly. The data obtained are integrated into a model for the structure of full-length YtvA, which presumably functions as a dimer.LOV2 domains (1), members of the superfamily of PAS domains (2, 3), are abundant in all domains of life and were first identified in plant phototropins (4). These photoreceptors regulate stomatal opening, phototropism, etc. and contain two N-terminal LOV domains that confer light regulation on the C-terminal Ser/Thr kinase domain (4). They also occur in bacteria, in which YtvA from Bacillus subtilis has been best characterized (for a review, see e.g. Ref. 5). Its N-terminal LOV domain binds FMN and shows the typical LOV photochemistry (6, 7): covalent adduct formation between a cysteine and the FMN chromophore. A linker helix, denoted Jα (7), connects the LOV domain to a STAS domain. The latter domain is present in many regulators of the general stress response of B. subtilis (8, 9). Stress via the addition of salt or ethanol (for a review, see Ref. 10) and blue light (11, 12) activates the general stress response via the environmental pathway, which integrates various signals via a large multiprotein complex, called the stressosome (13, 14). YtvA, which mediates light activation of σB (11, 12, 15), co-purifies with other STAS domain proteins in the stressosomes (16).When cells are stressed, STAS domains of several stressosome proteins (e.g. RsbS and RsbR) are phosphorylated by another intrinsic stressosome component, the serine/threonine kinase RsbT (9, 14, 17, 18). Next, RsbT is released from the complex to trigger RsbU, a protein phosphatase, thus (indirectly) activating σB (19). Phosphorylation of YtvA, however, has never been detected. Rather, it has been demonstrated in vitro that YtvA shows light-dependent GTP binding, presumably at its NTP-binding site in the STAS domain (20).Little is known about the mechanism of signal transmission in and by YtvA, except that in the C62A mutant, photochemistry in vitro (12) and light activation of σB in vivo (12, 15) are abolished. More detailed information is available for LOV domains of phototropins. A conserved glutamine, which is in hydrogen-bonding contact with the isoalloxazine ring of FMN, rotates its side chain by 180° upon covalent adduct formation (21). Replacement of this residue by leucine in the LOV2 domain of Phy3 from Adiantum results in a considerable reduction of the light-induced structural change (22). The corresponding mutation in phototropin 1 from Arabidopsis impairs autophosphorylation activity (23). The signal generated in the LOV2 domain is transmitted to the downstream kinase domain of phototropin 1 of Avena sativa through disruption of the interaction between its central β-sheet and the C-terminal linker region, the Jα-helix (24).Here, we study the mechanism of activation of YtvA in vivo, i.e. light-induced activation of the σB response, with site-directed mutagenesis. We focus on three regions of the protein, the flavin-binding pocket, the β-sheet of the LOV domain, and the GTP-binding site, and on potential phosphorylation sites of the STAS domain. We demonstrate that light-activated GTP binding is crucial for functional YtvA. A computational approach was used to model the structure of full-length YtvA. The model suggests that light modulates accessibility of the GTP-binding site of the STAS domain of YtvA.  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