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 390 毫秒
1.
A key set of reactions for the initiation of new DNA strands during herpes simplex virus-1 replication consists of the primase-catalyzed synthesis of short RNA primers followed by polymerase-catalyzed DNA synthesis (i.e. primase-coupled polymerase activity). Herpes primase (UL5-UL52-UL8) synthesizes products from 2 to ∼13 nucleotides long. However, the herpes polymerase (UL30 or UL30-UL42) only elongates those at least 8 nucleotides long. Surprisingly, coupled activity was remarkably inefficient, even considering only those primers at least 8 nucleotides long, and herpes polymerase typically elongated <2% of the primase-synthesized primers. Of those primers elongated, only 4–26% of the primers were passed directly from the primase to the polymerase (UL30-UL42) without dissociating into solution. Comparing RNA primer-templates and DNA primer-templates of identical sequence showed that herpes polymerase greatly preferred to elongate the DNA primer by 650–26,000-fold, thus accounting for the extremely low efficiency with which herpes polymerase elongated primase-synthesized primers. Curiously, one of the DNA polymerases of the host cell, polymerase α (p70-p180 or p49-p58-p70-p180 complex), extended herpes primase-synthesized RNA primers much more efficiently than the viral polymerase, raising the possibility that the viral polymerase may not be the only one involved in herpes DNA replication.Herpes simplex virus 1 (HSV-1)2 encodes seven proteins essential for replicating its double-stranded DNA genome; five of these encode the heterotrimeric helicase-primase (UL5-UL52-UL8 gene products) and the heterodimeric polymerase (UL30-UL42 gene products) (1, 2). The helicase-primase unwinds the DNA at the replication fork and generates single-stranded DNA for both leading and lagging strand synthesis. Primase synthesizes short RNA primers on the lagging strand that the polymerase presumably elongates using dNTPs (i.e. primase-coupled polymerase activity). These two protein complexes are thought to replicate the viral genome on both the leading and lagging strands (1, 2).Previous studies have focused on the helicase-primase and polymerase separately. The helicase-primase contains three subunits, UL5, UL52, and UL8 in a 1:1:1 ratio (35). The UL5 subunit has helicase-like motifs and the UL52 subunit has primase-like motifs, yet the minimal active complex that demonstrates either helicase or primase activities contains both UL5 and UL52 (6, 7). Although the UL8 subunit has no known catalytic activity, several functions have been proposed, including enhancing helicase and primase activities, enhancing primer synthesis on ICP8 (the HSV-1 single-stranded binding protein)-coated DNA strands, and facilitating formation of the replisome (812). Although primase will synthesize short (23 nucleotides long) primers on a variety of template sequences, synthesis of longer primers up to 13 nucleotides long requires the template sequence, 3′-deoxyguanidine-pyrimidine-pyrimidine-5′ (13). Primase initiates synthesis at the first pyrimidine via the polymerization of two purine NTPs (13). Even after initiation at this sequence, however, the vast majority of products are only 2–3 nucleotides long (13, 14).The herpes polymerase consists of the UL30 subunit, which has polymerase and 3′ → 5′ exonuclease activities (1, 2), and the UL42 subunit, which serves as a processivity factor (1517). Unlike most processivity factors that encircle the DNA, the UL42 protein binds double-stranded DNA and thus directly tethers the polymerase to the DNA (18). Using pre-existing DNA primer-templates as the substrate, the heterodimeric polymerase (UL30-UL42) incorporates dNTPs at a rate of 150 s–1, a rate much faster than primer synthesis (for primers >7 nucleotides long, 0.0002–0.01 s–1) (19, 20).We examined primase-coupled polymerase activity by the herpes primase and polymerase complexes. Although herpes primase synthesizes RNA primers 2–13 nucleotides long, the polymerase only effectively elongates those at least 8 nucleotides long. Surprisingly, the polymerase elongated only a small fraction of the primase-synthesized primers (<1–2%), likely because of the polymerase elongating RNA primer-templates much less efficiently than DNA primer-templates. In contrast, human DNA polymerase α (pol α) elongated the herpes primase-synthesized primers very efficiently. The biological significance of these data is discussed.  相似文献   

2.
Mcm10 is an essential eukaryotic protein required for the initiation and elongation phases of chromosomal replication. Specifically, Mcm10 is required for the association of several replication proteins, including DNA polymerase α (pol α), with chromatin. We showed previously that the internal (ID) and C-terminal (CTD) domains of Mcm10 physically interact with both single-stranded (ss) DNA and the catalytic p180 subunit of pol α. However, the mechanism by which Mcm10 interacts with pol α on and off DNA is unclear. As a first step toward understanding the structural details for these critical intermolecular interactions, x-ray crystallography and NMR spectroscopy were used to map the binary interfaces between Mcm10-ID, ssDNA, and p180. The crystal structure of an Mcm10-ID·ssDNA complex confirmed and extended our previous evidence that ssDNA binds within the oligonucleotide/oligosaccharide binding-fold cleft of Mcm10-ID. We show using NMR chemical shift perturbation and fluorescence spectroscopy that p180 also binds to the OB-fold and that ssDNA and p180 compete for binding to this motif. In addition, we map a minimal Mcm10 binding site on p180 to a small region within the p180 N-terminal domain (residues 286–310). These findings, together with data for DNA and p180 binding to an Mcm10 construct that contains both the ID and CTD, provide the first mechanistic insight into how Mcm10 might use a handoff mechanism to load and stabilize pol α within the replication fork.To maintain their genomic integrity, cells must ensure complete and accurate DNA replication once per cell cycle. Consequently, DNA replication is a highly regulated and orchestrated series of molecular events. Multiprotein complexes assembled at origins of replication lead to assembly of additional proteins that unwind chromosomal DNA and synthesize nascent strands. The first event is the formation of a pre-replicative complex, which is composed of the origin recognition complex, Cdc6, Cdt1, and Mcm2–7 (for review, see Ref. 1). Initiation of replication at the onset of S-phase involves the activity of cyclin- and Dbf4-dependent kinases concurrent with recruitment of key factors to the origin. Among these, Mcm10 (2, 3) is recruited in early S-phase and is required for loading of Cdc45 (4). Mcm2–7, Cdc45, and the GINS complex form the replicative helicase (58). Origin unwinding is followed by loading of RPA,3 And-1/Ctf4, and pol α onto ssDNA (912). In addition, recruitment of Sld2, Sld3, and Dpb11/TopBP1 are essential for replication initiation (13, 14), and association of topoisomerase I, proliferating cellular nuclear antigen (PCNA), replication factor C, and the replicative DNA polymerases δ and ϵ completes the replisome (for review, see Ref. 15).Mcm10 is exclusive to eukaryotes and is essential to both initiation and elongation phases of chromosomal DNA replication (6, 8, 16). Mutations in Mcm10 in yeast result in stalled replication, cell cycle arrest, and cell death (2, 3, 1719). These defects can be explained by the number of genetic and physical interactions between Mcm10 and many essential replication proteins, including origin recognition complex, Mcm2–7, and PCNA (3, 12, 2024). In addition, Mcm10 has been shown to stimulate the phosphorylation of Mcm2–7 by Dbf4-dependent kinase in vitro (25). Thus, Mcm10 is an integral component of the replication machinery.Importantly, Mcm10 physically interacts with and stabilizes pol α and helps to maintain its association with chromatin (16, 26, 27). This is a critical interaction during replication because pol α is the only enzyme in eukaryotic cells that is capable of initiating DNA synthesis de novo. Indeed, Mcm10 stimulates the polymerase activity of pol α in vitro (28), and interestingly, the fission yeast Mcm10, but not Xenopus Mcm10, has been shown to exhibit primase activity (29, 30). Mcm10 is composed of three domains, the N-terminal (NTD), internal (ID), and C-terminal (CTD) domains (29). The NTD is presumably an oligomerization domain, whereas the ID and CTD both interact with DNA and pol α (29). The CTD is not found in yeast, whereas the ID is highly conserved among all eukaryotes. The crystal structure of Mcm10-ID showed that this domain is composed of an oligonucleotide/oligosaccharide binding (OB)-fold and a zinc finger motif, which form a unified DNA binding platform (31). An Hsp10-like motif important for the interaction with pol α has been identified in the sequence of Saccharomyces cerevisiae Mcm10-ID (16, 26).DNA pol α-primase is composed of four subunits: p180, p68, p58, and p48. The p180 subunit possesses the catalytic DNA polymerase activity, and disruption of this gene is lethal (32, 33). p58 and p48 form the DNA-dependent RNA polymerase (primase) activity (34, 35), whereas the p68 subunit has no known catalytic activity but serves a regulatory role (36, 37). Pol α plays an essential role in lagging strand synthesis by first creating short (7–12 nucleotide) RNA primers followed by DNA extension. At the critical length of ∼30 nucleotides, replication factor C binds to the nascent strand to displace pol α and loads PCNA with pols δ and ϵ (for review, see Ref. 38).The interaction between Mcm10 and pol α has led to the suggestion that Mcm10 may help recruit the polymerase to the emerging replisome. However, the molecular details of this interaction and the mechanism by which Mcm10 may recruit and stabilize the pol α complex on DNA has not been investigated. Presented here is the high resolution structure of the conserved Mcm10-ID bound to ssDNA together with NMR chemical shift perturbation competition data for pol α binding in the presence of ssDNA. Collectively, these data demonstrate a shared binding site for DNA and pol α in the OB-fold cleft of Mcm10-ID, with a preference for ssDNA over pol α. In addition, we have mapped the Mcm10-ID binding site on pol α to a 24-residue segment of the N-terminal domain of p180. Based on these results, we propose Mcm10 helps to recruit pol α to origins of replication by a molecular hand-off mechanism.  相似文献   

3.
Integrin α8β1 interacts with a variety of Arg-Gly-Asp (RGD)-containing ligands in the extracellular matrix. Here, we examined the binding activities of α8β1 integrin toward a panel of RGD-containing ligands. Integrin α8β1 bound specifically to nephronectin with an apparent dissociation constant of 0.28 ± 0.01 nm, but showed only marginal affinities for fibronectin and other RGD-containing ligands. The high-affinity binding to α8β1 integrin was fully reproduced with a recombinant nephronectin fragment derived from the RGD-containing central “linker” segment. A series of deletion mutants of the recombinant fragment identified the LFEIFEIER sequence on the C-terminal side of the RGD motif as an auxiliary site required for high-affinity binding to α8β1 integrin. Alanine scanning mutagenesis within the LFEIFEIER sequence defined the EIE sequence as a critical motif ensuring the high-affinity integrin-ligand interaction. Although a synthetic LFEIFEIER peptide failed to inhibit the binding of α8β1 integrin to nephronectin, a longer peptide containing both the RGD motif and the LFEIFEIER sequence was strongly inhibitory, and was ∼2,000-fold more potent than a peptide containing only the RGD motif. Furthermore, trans-complementation assays using recombinant fragments containing either the RGD motif or LFEIFEIER sequence revealed a clear synergism in the binding to α8β1 integrin. Taken together, these results indicate that the specific high-affinity binding of nephronectin to α8β1 integrin is achieved by bipartite interaction of the integrin with the RGD motif and LFEIFEIER sequence, with the latter serving as a synergy site that greatly potentiates the RGD-driven integrin-ligand interaction but has only marginal activity to secure the interaction by itself.Integrins are a family of adhesion receptors that interact with a variety of extracellular ligands, typically cell-adhesive proteins in the extracellular matrix (ECM).2 They play mandatory roles in embryonic development and the maintenance of tissue architectures by providing essential links between cells and the ECM (1). Integrins are composed of two non-covalently associated subunits, termed α and β. In mammals, 18 α and 8 β subunits have been identified, and combinations of these subunits give rise to at least 24 distinct integrin heterodimers. Based on their ligand-binding specificities, ECM-binding integrins are classified into three groups, namely laminin-, collagen- and RGD-binding integrins (2, 3), of which the RGD-binding integrins have been most extensively investigated. The RGD-binding integrins include α5β1, α8β1, αIIbβ3, and αV-containing integrins, and have been shown to interact with a variety of ECM ligands, such as fibronectin and vitronectin, with distinct binding specificities.The α8 integrin subunit was originally identified in chick nerves (4). Integrin α8β1 is expressed in the metanephric mesenchyme and plays a crucial role in epithelial-mesenchymal interactions during the early stages of kidney morphogenesis. Disruption of the α8 gene in mice was found to be associated with severe defects in kidney morphogenesis (5) and stereocilia development (6). To date, α8β1 integrin has been shown to bind to fibronectin, vitronectin, osteopontin, latency-associated peptide of transforming growth factor-β1, tenascin-W, and nephronectin (also named POEM) (713), among which nephronectin is believed to be an α8β1 integrin ligand involved in kidney development (10).Nephronectin is one of the basement membrane proteins whose expression and localization patterns are restricted in a tissue-specific and developmentally regulated manner (10, 11). Nephronectin consists of five epidermal growth factor-like repeats, a linker segment containing the RGD cell-adhesive motif (designated RGD-linker) and a meprin-A5 protein-receptor protein-tyrosine phosphatase μ (MAM) domain (see Fig. 3A). Although the physiological functions of nephronectin remain only poorly understood, it is thought to play a role in epithelial-mesenchymal interactions through binding to α8β1 integrin, thereby transmitting signals from the epithelium to the mesenchyme across the basement membrane (10). Recently, mice deficient in nephronectin expression were produced by homologous recombination (14). These nephronectin-deficient mice frequently displayed kidney agenesis, a phenotype reminiscent of α8 integrin knock-out mice (14), despite the fact that other RGD-containing ligands, including fibronectin and osteopontin, were expressed in the embryonic kidneys (9, 15). The failure of the other RGD-containing ligands to compensate for the deficiency of nephronectin in the developing kidneys suggests that nephronectin is an indispensable α8β1 ligand that plays a mandatory role in epithelial-mesenchymal interactions during kidney development.Open in a separate windowFIGURE 3.Binding activities of α8β1 integrin to nephronectin and its fragments. A, schematic diagrams of full-length nephronectin (NN) and its fragments. RGD-linker and RGD-linker (GST), the central RGD-containing linker segments expressed in mammalian and bacterial expression systems, respectively; PRGDV, a short RGD-containing peptide modeled after nephronectin and expressed as a GST fusion protein (see Fig. 4A for the peptide sequence). The arrowheads indicate the positions of the RGD motif. B, purified recombinant proteins were analyzed by SDS-PAGE in 7–15% gradient (left and center panels) and 12% (right panels) gels, followed by Coomassie Brilliant Blue (CBB) staining, immunoblotting with an anti-FLAG mAb, or lectin blotting with PNA. The quantities of proteins loaded were: 0.5 μg (for Coomassie Brilliant Blue staining) and 0.1 μg (for blotting with anti-FLAG and PNA) in the left and center panels;1 μg in the right panel. C, recombinant proteins (10 nm) were coated on microtiter plates and assessed for their binding activities toward α8β1 integrin (10 nm) in the presence of 1 mm Mn2+. The backgrounds were subtracted as described in the legend to Fig. 2. The results represent the mean ± S.D. of triplicate determinations. D, titration curves of α8β1 integrin bound to full-length nephronectin (NN, closed squares), the RGD-linker segments expressed in 293F cells (RGD-linker, closed triangles) and E. coli (RGD-linker (GST), open triangles), the MAM domain (MAM, closed diamonds), and the PRGDV peptide expressed as a GST fusion protein in E. coli (PRGDV (GST), open circles). The assays were performed as described in the legend to Fig. 2B. The results represent the means of duplicate determinations.Although ligand recognition by RGD-binding integrins is primarily determined by the RGD motif in the ligands, it is the residues outside the RGD motif that define the binding specificities and affinities toward individual integrins (16, 17). For example, α5β1 integrin specifically binds to fibronectin among the many RGD-containing ligands, and requires not only the RGD motif in the 10th type III repeat but also the so-called “synergy site” within the preceding 9th type III repeat for fibronectin recognition (18). Recently, DiCara et al. (19) demonstrated that the high-affinity binding of αVβ6 integrin to its natural ligands, e.g. foot-and-mouth disease virus, requires the RGD motif immediately followed by a Leu-Xaa-Xaa-Leu/Ile sequence, which forms a helix to align the two conserved hydrophobic residues along the length of the helix. Given the presence of many naturally occurring RGD-containing ligands, it is conceivable that the specificities of the RGD-binding integrins are dictated by the sequences flanking the RGD motif or those in neighboring domains that come into close proximity with the RGD motif in the intact ligand proteins. However, the preferences of α8β1 integrin for RGD-containing ligands and how it secures its high-affinity binding toward its preferred ligands remain unknown.In the present study, we investigated the binding specificities of α8β1 integrin toward a panel of RGD-containing cell-adhesive proteins. Our data reveal that nephronectin is a preferred ligand for α8β1 integrin, and that a LFEIFEIER sequence on the C-terminal side of its RGD motif serves as a synergy site to ensure the specific high-affinity binding of nephronectin to α8β1 integrin.  相似文献   

4.
5.
Studies of replicative DNA polymerases have led to the generalization that abasic sites are strong blocks to DNA replication. Here we show that yeast replicative DNA polymerase ϵ bypasses a model abasic site with comparable efficiency to Pol η and Dpo4, two translesion polymerases. DNA polymerase ϵ also exhibited high bypass efficiency with a natural abasic site on the template. Translesion synthesis primarily resulted in deletions. In cases where only a single nucleotide was inserted, dATP was the preferred nucleotide opposite the natural abasic site. In contrast to translesion polymerases, DNA polymerase ϵ with 3′–5′ proofreading exonuclease activity bypasses only the model abasic site during processive synthesis and cannot reinitiate DNA synthesis. This characteristic may allow other pathways to rescue leading strand synthesis when stalled at an abasic site.  相似文献   

6.
The nuclear-encoded DNA polymerase γ (DNA POLγ) is the sole DNA polymerase required for the replication of the mitochondrial DNA. We have cloned the cDNA for human DNA POLγ and have mapped the gene to the chromosomal location 15q24. Additionally, the DNA POLγ gene fromDrosophila melanogasterand a partial cDNA for DNA POLγ fromGallus gallushave been cloned. The predicted human DNA POLγ polypeptide is 1239 amino acids, with a calculated molecular mass of 139.5 kDa. The human amino acid sequence is 41.6, 43.0, 48.7, and 77.6% identical to those ofSchizosaccharomyces pombe, Saccharomyces cerevisiae, Drosophila melanogaster,and the C-terminal half ofG. gallus,respectively. Polyclonal antibodies raised against the polymerase portion of the protein reacted specifically with a 140-kDa protein in mitochondrial extracts and immunoprecipitated a protein with DNA POLγ like activity from mitochondrial extracts. The human DNA POLγ is unique in that the first exon of the gene contains a CAG10trinucleotide repeat.  相似文献   

7.
8.
Recently we reported that N-glycans on the β-propeller domain of the integrin α5 subunit (S-3,4,5) are essential for α5β1 heterodimerization, expression, and cell adhesion. Herein to further investigate which N-glycosylation site is the most important for the biological function and regulation, we characterized the S-3,4,5 mutants in detail. We found that site-4 is a key site that can be specifically modified by N-acetylglucosaminyltransferase III (GnT-III). The introduction of bisecting GlcNAc into the S-3,4,5 mutant catalyzed by GnT-III decreased cell adhesion and migration on fibronectin, whereas overexpression of N-acetylglucosaminyltransferase V (GnT-V) promoted cell migration. The phenomenon is similar to previous observations that the functions of the wild-type α5 subunit were positively and negatively regulated by GnT-V and GnT-III, respectively, suggesting that the α5 subunit could be duplicated by the S-3,4,5 mutant. Interestingly GnT-III specifically modified the S-4,5 mutant but not the S-3,5 mutant. This result was confirmed by erythroagglutinating phytohemagglutinin lectin blot analysis. The reduction in cell adhesion was consistently observed in the S-4,5 mutant but not in the S-3,5 mutant cells. Furthermore mutation of site-4 alone resulted in a substantial decrease in erythroagglutinating phytohemagglutinin lectin staining and suppression of cell spread induced by GnT-III compared with that of either the site-3 single mutant or wild-type α5. These results, taken together, strongly suggest that N-glycosylation of site-4 on the α5 subunit is the most important site for its biological functions. To our knowledge, this is the first demonstration that site-specific modification of N-glycans by a glycosyltransferase results in functional regulation.Glycosylation is a crucial post-translational modification of most secreted and cell surface proteins (1). Glycosylation is involved in a variety of physiological and pathological events, including cell growth, migration, differentiation, and tumor invasion. It is well known that glycans play important roles in cell-cell communication, intracellular signal transduction, protein folding, and stability (2, 3).Integrins comprise a family of receptors that are important for cell adhesion. The major function of integrins is to connect cells to the extracellular matrix, activate intracellular signaling pathways, and regulate cytoskeletal formation (4). Integrin α5β1 is well known as a fibronectin (FN)3 receptor. The interaction between integrin α5 and FN is essential for cell migration, cell survival, and development (58). In addition, integrins are N-glycan carrier proteins. For example, α5β1 integrin contains 14 and 12 putative N-glycosylation sites on the α5 and β1 subunits, respectively. Several studies suggest that N-glycosylation is essential for functional integrin α5β1. When human fibroblasts were cultured in the presence of 1-deoxymannojirimycin, which prevents N-linked oligosaccharide processing, immature α5β1 integrin appeared on the cell surface, and FN-dependent adhesion was greatly reduced (9). Treatment of purified integrin α5β1 with N-glycosidase F, which cleaves between the innermost N-acetylglucosamine (GlcNAc) and asparagine N-glycan residues of N-linked glycoproteins, prevented the inherent association between subunits and blocked α5β1 binding to FN (10).A growing body of evidence indicates that the presence of the appropriate oligosaccharide can modulate integrin activation. N-Acetylglucosaminyltransferase III (GnT-III) catalyzes the addition of GlcNAc to mannose that is β1,4-linked to an underlying N-acetylglucosamine, producing what is known as a “bisecting” GlcNAc linkage as shown in Fig. 1B. GnT-III is generally regarded as a key glycosyltransferase in N-glycan biosynthetic pathways and contributes to inhibition of metastasis. The introduction of a bisecting GlcNAc catalyzed by GnT-III suppresses additional processing and elongation of N-glycans. These reactions, which are catalyzed in vitro by other glycosyltransferases, such as N-acetylglucosaminyltransferase V (GnT-V), which catalyzes the formation of β1,6 GlcNAc branching structures (Fig. 1B) and plays important roles in tumor metastasis, do not proceed because the enzymes cannot utilize the bisected N-glycans as a substrate. Introduction of the bisecting GlcNAc to integrin α5 by overexpression of GnT-III resulted in decreased in ligand binding and down-regulation of cell adhesion and migration (1113). Contrary to the functions of GnT-III, overexpression of GnT-V promoted integrin α5β1-mediated cell migration on FN (14). These observations clearly demonstrate that the alteration of N-glycan structure affected the biological functions of integrin α5β1. Similarly characterization of the carbohydrate moieties in integrin α3β1 from non-metastatic and metastatic human melanoma cell lines showed that expression of β1,6 GlcNAc branched structures was higher in metastatic cells compared with non-metastatic cells, confirming the notion that the β1,6 GlcNAc branched structure confers invasive and metastatic properties to cancer cells. In fact, Partridge et al. (15) reported that GnT-V-modified N-glycans containing poly-N-acetyllactosamine, the preferred ligand for galectin-3, on surface receptors oppose their constitutive endocytosis, promoting intracellular signaling and consequently cell migration and tumor metastasis.Open in a separate windowFIGURE 1.Potential N-glycosylation sites on the α5 subunit and its modification by GnT-III and GnT-V. A, schematic diagram of potential N-glycosylation sites on the α5 subunit. Putative N-glycosylation sites are indicated by triangles, and point mutations are indicated by crosses (N84Q, N182Q, N297Q, N307Q, N316Q, N524Q, N530Q, N593Q, N609Q, N675Q, N712Q, N724Q, N773Q, and N868Q). B, illustration of the reaction catalyzed by GnT-III and GnT-V. Square, GlcNAc; circle, mannose. TM, transmembrane domain.In addition, sialylation on the non-reducing terminus of N-glycans of α5β1 integrin plays an important role in cell adhesion. Colon adenocarcinomas express elevated levels of α2,6 sialylation and increased activity of ST6GalI sialyltransferase. Elevated ST6GalI positively correlated with metastasis and poor survival. Therefore, ST6GalI-mediated hypersialylation likely plays a role in colorectal tumor invasion (16, 17). In fact, oncogenic ras up-regulated ST6GalI and, in turn, increased sialylation of β1 integrin adhesion receptors in colon epithelial cells (18). However, this is not always the case. The expression of hyposialylated integrin α5β1 was induced by phorbol esterstimulated differentiation in myeloid cells in which the expression of the ST6GalI was down-regulated by the treatment, increasing FN binding (19). A similar phenomenon was also observed in hematopoietic or other epithelial cells. In these cells, the increased sialylation of the β1 integrin subunit was correlated with reduced adhesiveness and metastatic potential (2022). In contrast, the enzymatic removal of α2,8-linked oligosialic acids from the α5 integrin subunit inhibited cell adhesion to FN (23). Collectively these findings suggest that the interaction of integrin α5β1 with FN is dependent on its N-glycosylation and the processing status of N-glycans.Because integrin α5β1 contains multipotential N-glycosylation sites, it is important to determine the sites that are crucial for its biological function and regulation. Recently we found that N-glycans on the β-propeller domain (sites 3, 4, and 5) of the integrin α5 subunit are essential for α5β1 heterodimerization, cell surface expression, and biological function (24). In this study, to further investigate the underlying molecular mechanism of GnT-III-regulated biological functions, we characterized the N-glycans on the α5 subunit in detail using genetic and biochemical approaches and found that site-4 is a key site that can be specifically modified by GnT-III.  相似文献   

9.
TANK-binding kinase 1 (TBK1) is a key component of Toll-like receptor-dependent and -independent signaling pathways. In response to microbial components, TBK1 activates interferon regulatory factor 3 (IRF3) and cytokine expression. Here we show that TBK1 is a novel target of the γ134.5 protein, a virulence factor whose expression is regulated in a temporal fashion. Remarkably, the γ134.5 protein is required to inhibit IRF3 phosphorylation, nuclear translocation, and the induction of antiviral genes in infected cells. When expressed in mammalian cells, the γ134.5 protein forms complexes with TBK1 and disrupts the interaction of TBK1 and IRF3, which prevents the induction of interferon and interferon-stimulated gene promoters. Down-regulation of TBK1 requires the amino-terminal domain. In addition, unlike wild type virus, a herpes simplex virus mutant lacking γ134.5 replicates efficiently in TBK1-/- cells but not in TBK1+/+ cells. Addition of exogenous interferon restores the antiviral activity in both TBK1-/- and TBK+/+ cells. Hence, control of TBK1-mediated cell signaling by the γ134.5 protein contributes to herpes simplex virus infection. These results reveal that TBK1 plays a pivotal role in limiting replication of a DNA virus.Herpes simplex virus 1 (HSV-1)3 is a large DNA virus that establishes latent or lytic infection, in which the virus triggers innate immune responses. In HSV-infected cells, a number of antiviral mechanisms operate in a cell type- and time-dependent manner (1). In response to double-stranded RNA (dsRNA), Toll-like receptor 3 (TLR3) recruits an adaptor TIR domain-containing adaptor inducing IFN-β and stimulates cytokine expression (2, 3). In the cytoplasm, RNA helicases, RIG-I (retinoid acid-inducible gene-I), and MDA5 (melanoma differentiation associated gene 5) recognize intracellular viral 5′-triphosphate RNA or dsRNA (2, 4). Furthermore, a DNA-dependent activator of IFN-regulatory factor (DAI) senses double-stranded DNA in the cytoplasm and induces cytokine expression (5). There is also evidence that viral entry induces antiviral programs independent of TLR and RIG-I pathways (6). While recognizing distinct viral components, these innate immune pathways relay signals to the two IKK-related kinases, TANK-binding kinase 1 (TBK1) and inducible IκB kinase (IKKi) (2).The IKK-related kinases function as essential components that phosphorylate IRF3 (interferon regulatory factor 3), as well as the closely related IRF7, which translocates to the nucleus and induces antiviral genes, such as interferon-α/β and ISG56 (interferon-stimulated gene 56) (7, 8). TBK1 is constitutively expressed, whereas IKKi is engaged as an inducible gene product of innate immune signaling (9, 10). IRF3 activation is attenuated in TBK1-deficient but not in IKKi-deficient cells (11, 12). Its activation is completely abolished in double-deficient cells (12), suggesting a partially redundant function of TBK1 and IKKi. Indeed, IKKi also negatively regulates the STAT-signaling pathway (13). TBK1/IKKi interacts with several proteins, such as TRAF family member-associated NF-κB activator (TANK), NAP1 (NAK-associated protein 1), similar to NAP1TBK1 adaptor (SINTBAD), DNA-dependent activator of IFN-regulatory factors (DAI), and secretory protein 5 (Sec5) in host cells (5, 1418). These interactions are thought to regulate TBK1/IKKi, which delineates innate as well as adaptive immune responses.Upon viral infection, expression of HSV proteins interferes with the induction of antiviral immunity. When treated with UV or cycloheximide, HSV induces an array of antiviral genes in human lung fibroblasts (19, 20). Furthermore, an HSV mutant, with deletion in immediate early protein ICP0, induces ISG56 expression (21). Accordingly, expression of ICP0 inhibits the induction of antiviral programs mediated by IRF3 or IRF7 (2123). However, although ICP0 negatively regulates IFN-β expression, it is not essential for this effect (24). In HSV-infected human macrophages or dendritic cells, an immediate early protein ICP27 is required to suppress cytokine induction involving IRF3 (25). In this context, it is notable that an HSV mutant, lacking a leaky late gene γ134.5, replicates efficiently in cells devoid of IFN-α/β genes (26). Additionally, the γ134.5 null mutant induces differential cytokine expression as compared with wild type virus (27). Thus, HSV modulation of cytokine expression is a complex process that involves multiple viral components. Currently, the molecular mechanism governing this event is unclear. In this study, we show that HSV γ134.5 targets TBK1 and inhibits antiviral signaling. The data herein reveal a previously unrecognized mechanism by which γ134.5 facilitates HSV replication.  相似文献   

10.
The interaction between epithelial cells and the extracellular matrix is crucial for tissue architecture and function and is compromised during cancer progression. Dystroglycan is a membrane receptor that mediates interactions between cells and basement membranes in various epithelia. In many epithelium-derived cancers, β-dystroglycan is expressed, but α-dystroglycan is not detected. Here we report that α-dystroglycan is correctly expressed and trafficked to the cell membrane but lacks laminin binding as a result of the silencing of the like-acetylglucosaminyltransferase (LARGE) gene in a cohort of highly metastatic epithelial cell lines derived from breast, cervical, and lung cancers. Exogenous expression of LARGE in these cancer cells restores the normal glycosylation and laminin binding of α-dystroglycan, leading to enhanced cell adhesion and reduced cell migration in vitro. Our findings demonstrate that LARGE repression is responsible for the defects in dystroglycan-mediated cell adhesion that are observed in epithelium-derived cancer cells and point to a defect of dystroglycan glycosylation as a factor in cancer progression.Normal epithelial cells are tightly associated with one another and with the underlying basement membrane to maintain tissue architecture and function. During cancer progression, primitive cancer cells escape from this control by modifying the binding affinities of their cell membrane receptors. Several receptors have been described as important for this process. Of these, the integrins are the best studied (1). The receptor dystroglycan has been reported to be required for the development and maintenance of epithelial tissues (2, 3). A direct requirement for dystroglycan in epithelia is further demonstrated by the profound effect that loss of dystroglycan expression has on cell polarity and laminin binding in cultured mammary epithelial cells (4, 5). However, dystroglycan is not only important in the establishment and maintenance of epithelial structure. Associations have also been made between the loss of α-dystroglycan immunoreactivity and cancer progression in tumors of epithelial origin, including breast, colon, cervix, and prostate cancers (4, 69). The dystroglycan loss of function could thus serve as an effective means by which cancerous cells modify their adhesion to the extracellular matrix (ECM).2Dystroglycan is a ubiquitously expressed cell membrane protein that plays a key function in cellular integrity, linking the intracellular cytoskeleton to the extracellular matrix. The dystroglycan gene encodes a preprotein that is cleaved into two peptides (10). The C-terminal component, known as β-dystroglycan, is embedded within the cell membrane, whereas the N-terminal component, α-dystroglycan, is present within the extracellular periphery but remains associated with β-dystroglycan through non-covalent bonds. β-Dystroglycan binds to actin (11), dystrophin (11), utrophin (11), and Grb2 (12) through its C-terminal intracellular domain. α-Dystroglycan, on the other hand, binds to ECM proteins that contain laminin globular domains including laminins (13, 14), agrin (15), and perlecan (16), as well as to the transmembrane protein neurexin (17). α-Dystroglycan is extensively decorated by three different types of glycan modifications: mucin type O-glycosylation, O-mannosylation, and N-glycosylation. The state of α-dystroglycan glycosylation has been shown to be critical for the ability of the protein to bind to laminin globular domain-containing proteins of the ECM (18).Previous studies of epithelium-derived cancers (4, 9) demonstrated that the loss of immunoreactivity of α-dystroglycan antibodies correlates with tumor grade and poor prognosis. This reduced detection of α-dystroglycan, however, is based on a loss of α-dystroglycan reactivity to antibodies (known as IIH6 and VIA4-1) that recognize the laminin-binding glyco-epitope of α-dystroglycan, i.e. the protein is only functional when it is glycosylated in such a way (henceforth, referred to as functional glycosylation). However, in most of the cancer samples that have been studied to date, β-dystroglycan is expressed at normal levels at the cell membrane. Thus, the aforementioned cancer-associated loss of α-dystroglycan expression may reflect a failure in the post-translational processing of dystroglycan rather than in the synthesis of α-dystroglycan itself.A similar defect in dystroglycan has been reported in a group of congenital muscular dystrophies (19). This spectrum of human developmental syndromes involves the brain, eye, and skeletal muscle and shows a dramatic gradient of phenotypic severity that ranges from the most devastating in Walker-Warburg syndrome to the least severe in limb-girdle muscular dystrophy. Six distinct known and putative glycosyltransferases have been shown to underlie these syndromes: protein O-mannosyltransferase 1 (POMT1), protein O-mannosyltransferase 2 (POMT2), protein O-mannose β-1,2-acetylglucosaminyltransferase 1 (POMGnT1), like acetylglucosaminyltransferase (LARGE), Fukutin, and Fukutin-related protein (FKRP) (2025). Indeed, all muscular dystrophy patients with mutations in any of these genes fail to express the functionally glycosylated α-dystroglycan epitope that is recognized by the IIH6 and VIA4-1 antibodies.To investigate the molecular mechanism responsible for the loss of α-dystroglycan in epithelium-derived cancers and its role in metastatic progression, we examined the expression and glycosylation status of α-dystroglycan in a group of breast, cervical, and lung cancer cell lines. Here we report that although α-dystroglycan is expressed in the metastatic cell lines MDA-MB-231, HeLa, H1299, and H2030, it is not functionally glycosylated. In screening these cell lines for expression of the six known α-dystroglycan-modifying proteins, we observed that only one, LARGE, was extensively down-regulated. We also report that the ectopic restoration of LARGE expression in these cell lines led not only to the production of a functional dystroglycan but also to the reversion of certain characteristics associated with invasiveness, namely cell attachment to ECM proteins and cell migration.  相似文献   

11.
12.
Apurinic/apyrimidinic (AP) sites are continuously generated in genomic DNA. Left unrepaired, AP sites represent noninstructional premutagenic lesions that are impediments to DNA synthesis. When DNA polymerases encounter an AP site, they generally insert dAMP. This preferential insertion is referred to as the A-rule. Crystallographic structures of DNA polymerase (pol) β, a family X polymerase, with active site mismatched nascent base pairs indicate that the templating (i.e. coding) base is repositioned outside of the template binding pocket thereby diminishing interactions with the incorrect incoming nucleotide. This effectively produces an abasic site because the template pocket is devoid of an instructional base. However, the template pocket is not empty; an arginine residue (Arg-283) occupies the space vacated by the templating nucleotide. In this study, we analyze the kinetics of pol β insertion opposite an AP site and show that the preferential incorporation of dAMP is lost with the R283A mutant. The crystallographic structures of pol β bound to gapped DNA with an AP site analog (tertrahydrofuran) in the gap (binary complex) and with an incoming nonhydrolyzable dATP analog (ternary complex) were solved. These structures reveal that binding of the dATP analog induces a closed polymerase conformation, an unstable primer terminus, and an upstream shift of the templating residue even in the absence of a template base. Thus, dATP insertion opposite an abasic site and dATP misinsertions have common features.  相似文献   

13.
The alpha subunit of the avian myeloblastosis virus DNA polymerase could be readily purified to near homogeneity using a polyuridylic acid-Sepharose column chromatography step.  相似文献   

14.
The activity of the mitochondrial replicase, DNA polymerase γ (Pol γ) is stimulated by another key component of the mitochondrial replisome, the mitochondrial single-stranded DNA-binding protein (mtSSB). We have performed a comparative analysis of the human and Drosophila Pols γ with their cognate mtSSBs, evaluating their functional relationships using a combined approach of biochemical assays and electron microscopy. We found that increasing concentrations of both mtSSBs led to the elimination of template secondary structure and gradual opening of the template DNA, through a series of visually similar template species. The stimulatory effect of mtSSB on Pol γ on these ssDNA templates is not species-specific. We observed that human mtSSB can be substituted by its Drosophila homologue, and vice versa, finding that a lower concentration of insect mtSSB promotes efficient stimulation of either Pol. Notably, distinct phases of the stimulation by both mtSSBs are distinguishable, and they are characterized by a similar organization of the template DNA for both Pols γ. We conclude that organization of the template DNA is the major factor contributing to the stimulation of Pol γ activity. Additionally, we observed that human Pol γ preferentially utilizes compacted templates, whereas the insect enzyme achieves its maximal activity on open templates, emphasizing the relative importance of template DNA organization in modulating Pol γ activity and the variation among systems.  相似文献   

15.
The transient protein-protein interactions induced by guanine nucleotide-dependent conformational changes of G proteins play central roles in G protein-coupled receptor-mediated signaling systems. Leukemia-associated RhoGEF (LARG), a guanine nucleotide exchange factor for Rho, contains an RGS homology (RH) domain and Dbl homology/pleckstrin homology (DH/PH) domains and acts both as a GTPase-activating protein (GAP) and an effector for Gα13. However, the molecular mechanism of LARG activation upon Gα13 binding is not yet well understood. In this study, we analyzed the Gα13-LARG interaction using cellular and biochemical methods, including a surface plasmon resonance (SPR) analysis. The results obtained using various LARG fragments demonstrated that active Gα13 interacts with LARG through the RH domain, DH/PH domains, and C-terminal region. However, an alanine substitution at the RH domain contact position in Gα13 resulted in a large decrease in affinity. Thermodynamic analysis revealed that binding of Gα13 proceeds with a large negative heat capacity change (ΔCp°), accompanied by a positive entropy change (ΔS°). These results likely indicate that the binding of Gα13 with the RH domain triggers conformational rearrangements between Gα13 and LARG burying an exposed hydrophobic surface to create a large complementary interface, which facilitates complex formation through both GAP and effector interfaces, and activates the RhoGEF. We propose that LARG activation is regulated by an induced-fit mechanism through the GAP interface of Gα13.Heterotrimeric G proteins3 serve as key molecular switches to transduce a large array of extracellular signals into cells by actively alternating their conformations between GDP-bound inactive and GTP-bound active forms. In the current model, the ligand-activated G protein-coupled receptors (GPCRs) catalyze the exchange of GDP for GTP on Gα subunits (1). Upon activation, three switch regions in the Gα subunit undergo significant conformational changes, followed by dissociation of the GTP-bound Gα subunit from the Gβγ subunits. Both Gα-GTP and free Gβγ interact with diverse downstream effectors to transmit intracellular signals. The Gα subunit hydrolyzes bound GTP to GDP by its intrinsic GTPase activity. This deactivation process is further accelerated by GTPase-activating proteins (GAPs) such as regulator of G protein signaling (RGS) proteins (2, 3). Gα-GDP dissociates from effectors and re-associates with Gβγ to terminate the signal.Although this model explains the basic concept of G protein signaling, the molecular dynamics of interactions among GPCR, G protein, RGS protein, and effector during the signaling process is not well understood. It has been suggested that the GPCR signals are integrated into the intracellular signaling network at the level of G proteins (4). Accumulating evidence suggests that the Gα subunit acts as the core of the signaling complex at the membrane, which is formed through the transient protein-protein interactions of multiple signaling components (5, 6). Thus, the quantitative analysis of the dynamic molecular interactions in the GPCR signaling complex will be crucial to understanding various cellular processes.Gα12 and Gα13 subunits have been demonstrated to regulate the activity of Rho GTPase through RhoGEFs, which contain an N-terminal RGS homology domain (RH-RhoGEFs) (710). RH-RhoGEFs, which consist of p115RhoGEF/Lsc, PDZ-Rho-GEF/GTRAP48, and LARG in mammalian species, directly link the activation of GPCRs by extracellular ligands to the regulation of Rho activity in cells (1014). All three RH-RhoGEFs contain an N-terminal RH domain, which specifically recognizes the active form of Gα12 or Gα13 and central DH/PH domains characteristic of GEFs for Rho GTPases. It has been demonstrated in vitro that LARG and p115RhoGEF serve as specific GAPs for Gα12/13 through their RH domains and also as their effectors to regulate Rho GTPase activation (1113). A structural study has demonstrated that the interface of the RH domain of p115RhoGEFs and a Gα13/i1 chimera is different from that of the RGS domain of RGS4 and Gαi1 (7). The N-terminal small element in the RH domain, which is required for GAP activity toward Gα13, contacts the switch regions and the helical domain of the Gα13/i1 chimera. The core module of the p115RhoGEF RH domain binds to the region of Gα13/i1, which is conventionally used for effector binding. These results suggest roles for the RH domain in the stimulation of GEF activity by Gα13 in addition to GAP activity. On the other hand, several studies have also indicated that regions outside of RH domain of RH-RhoGEFs, particularly the DH/PH domains, interact directly with activated Gα13 (11, 14, 15). In addition, we have demonstrated recently that p115RhoGEF interacts with distinct surfaces of Gα13 for the GAP reaction or GEF activity regulation (16). However, the molecular mechanism of LARG activation upon Gα13 binding is not clearly understood.In this study, we have developed a quantitative method for the kinetic and thermodynamic analysis of Gα13-effector interaction using surface plasmon resonance (SPR) with sensor chips on which Gα13 was immobilized. We examined the kinetics and thermodynamics of the Gα13-LARG interaction and assessed LARG activation using both in vitro and cell-based approaches. We present evidence that, in addition to the interaction with the RH domain, the DH/PH domains and C-terminal region of LARG also interact with Gα13 to form the high affinity Gα13-LARG complex and activate RhoGEF activity. We further propose that LARG adopts the active conformation using an induced-fit mechanism through association with the GAP interface of Gα13. A similar mechanism may also be used with other Gα-effector interactions.  相似文献   

16.
We have previously shown that the C-terminal cystathionine β-synthase (CBS) domains of the nucleotide-binding domains of the ABC transporter OpuA, in conjunction with an anionic membrane surface function, act as sensor of internal ionic strength (Iin). Here, we show that a surface-exposed cationic region in the CBS module domain is critical for ion sensing. The consecutive substitution of up to five cationic residues led to a gradual decrease of the ionic strength dependence of transport. In fact, a 5-fold mutant was essentially independent of salt in the range from 0 to 250 mm KCl (or NaCl), supplemented to medium of 30 mm potassium phosphate. Importantly, the threshold temperature for transport was lowered by 5–7 °C and the temperature coefficient Q10 was lowered from 8 to ∼1.5 in the 5-fold mutant, indicating that large conformational changes are accompanying the CBS-mediated regulation of transport. Furthermore, by replacing the anionic C-terminal tail residues that extend the CBS module with histidines, the transport of OpuA became pH-dependent, presumably by additional charge interactions of the histidine residues with the membrane. The pH dependence was not observed at high ionic strength. Altogether the analyses of the CBS mutants support the notion that the osmotic regulation of OpuA involves a simple biophysical switching mechanism, in which nonspecific electrostatic interactions of a protein module with the membrane are sufficient to lock the transporter in the inactive state.In their natural habitats microorganisms are often exposed to changes in the concentration of solutes in the environment (1). A sudden increase in the medium osmolality results in loss of water from the cell, loss of turgor, a decrease in cell volume, and an increase in intracellular osmolyte concentration. Osmoregulatory transporters such as OpuA in Lactococcus lactis, ProP in Escherichia coli, and BetP in Corynebacterium glutamicum diminish the consequences of the osmotic stress by mediating the uptake of compatible solutes upon an increase in extracellular osmolality (24). For the ATP-binding cassette (ABC)5 transporter OpuA, it has been shown that the system, reconstituted in proteoliposomes, is activated by increased concentrations of lumenal ions (increased internal ionic strength) (2, 5, 6). This activation is instantaneous both in vivo and in vitro and only requires threshold levels of ionic osmolytes. Moreover, the ionic threshold for activation is highly dependent of the ionic lipid content (charge density) of the membrane and requires the presence of so-called cystathionine β-synthase (CBS) domains, suggesting that the ionic signal is transduced to the transporter via critical interactions of the protein with membrane lipids.The ABC transporter OpuA consists of two identical nucleotide-binding domains (NBD) fused to CBS domains and two identical substrate-binding domains fused to transmembrane domains. The NBD-CBS and substrate-binding domain-transmembrane domain subunits are named OpuAA and OpuABC, respectively. Two tandem CBS domains are linked to the C-terminal end of the NBD; each domain (CBS1 and CBS2) has a β-α-β-β-α secondary structure (5) (Fig. 1A). The CBS domains are widely distributed in most if not all species of life but their function is largely unknown. Most of the CBS domains are found as tandem repeats but data base searches have also revealed tetra-repeat units (5). The crystal structures of several tandem CBS domains have been elucidated (79, 32), and in a number of cases it has been shown that two tandem CBS domains form dimeric structures with a total of four CBS domains per structural module (hereafter referred to as CBS module). The crystal structures of the full-length MgtE Mg2+ transporter confirm the dimeric configuration and show that the CBS domains undergo large conformational changes upon Mg2+ binding or release (10, 11). In general, ABC transporters are functional as dimers, which implies that two tandem CBS domains are present in the OpuA complex. Preliminary experiments with disulfides engineered at the interface of two tandem CBS domains in OpuA suggest that large structural rearrangements (association-dissociation of the interfaces) play a determining role in the ionic strength-regulated transport. Finally, a subset of CBS-containing proteins has a C-terminal extension, which in OpuA is highly anionic (sequence: ADIPDEDEVEEIEKEEENK) and modulates the ion sensing activity (6).Open in a separate windowFIGURE 1.Domain structure of CBS module of OpuA. A, sequence of tandem CBS domains. The predicted secondary structure is indicated above the sequence. The residues modified in this study are underlined. The amino acid sequence end-points of OpuAΔ61 and OpuAΔ119 are indicated by vertical arrows. B, homology model of tandem CBS domain of OpuA. The CBS domains were individually modeled on the crystal structure of the tandem CBS protein Ta0289 from T. acidophilum (PDB entry 1PVM), using Phyre. Ta0289 was used for the initial modeling, because its primary sequence was more similar to the CBS domains of OpuA than those of the other crystallized CBS proteins. The individual domain models were then assembled with reference to the atomic coordinates of the tandem CBS domains of IMPDH from Streptococcus pyogenes (PDB entry 1ZFJ) to form the tandem CBS pair, using PyMOL (DeLano). The positions of the (substituted) cationic residues are indicated.In this study, we have engineered the surface-exposed cationic residues of the CBS module and the C-terminal anionic tail of OpuA (Fig. 1B). The ionic strength and lipid dependence of the OpuA mutants were determined in vivo and in vitro. We show that substitution of five cationic residues for neutral amino acids is sufficient to inactivate the ionic strength sensor and convert OpuA into a constitutively active transporter. Moreover, by substituting six anionic plus four neutral residues of the C-terminal anionic tail for histidines, the transport reaction becomes strongly pH-dependent.  相似文献   

17.
18.
Human polymerase kappa (hPol κ) is one of four eukaryotic Y-class DNA polymerases and may be an important element in the cellular response to polycyclic aromatic hydrocarbons such as benzo[a]pyrene, which can lead to reactive oxygenated metabolite-mediated oxidative stress. Here, we present a detailed analysis of the activity and specificity of hPol κ bypass opposite the major oxidative adduct 7,8-dihydro-8-oxo-2′-deoxyguanosine (8-oxoG). Unlike its archaeal homolog Dpo4, hPol κ bypasses this lesion in an error-prone fashion by inserting mainly dATP. Analysis of transient-state kinetics shows diminished “bursts” for dATP:8-oxoG and dCTP:8-oxoG incorporation, indicative of non-productive complex formation, but dATP:8-oxoG insertion events that do occur are 2-fold more efficient than dCTP:G insertion events. Crystal structures of ternary hPol κ complexes with adducted template-primer DNA reveal non-productive (dGTP and dATP) alignments of incoming nucleotide and 8-oxoG. Structural limitations placed upon the hPol κ by interactions between the N-clasp and finger domains combined with stabilization of the syn-oriented template 8-oxoG through the side chain of Met-135 both appear to contribute to error-prone bypass. Mutating Leu-508 in the little finger domain of hPol κ to lysine modulates the insertion opposite 8-oxoG toward more accurate bypass, similar to previous findings with Dpo4. Our structural and activity data provide insight into important mechanistic aspects of error-prone bypass of 8-oxoG by hPol κ compared with accurate and efficient bypass of the lesion by Dpo4 and polymerase η.DNA damage incurred by a multitude of endogenous and exogenous factors constitutes an inevitable challenge for the replication machinery, and various mechanisms exist to either remove the resulting lesions or bypass them in a more or less mutation-prone fashion (1). Error-prone polymerases are central to trans-lesion synthesis across sites of damaged DNA (2, 3). Four so-called Y-class DNA polymerases have been identified in humans, Pol η,4 Pol ι, Pol κ, and Rev1, which exhibit different activities and abilities to replicate past a flurry of individual lesions (4, 5). Homologs have also been identified and characterized in other organisms, notably DinB (Pol IV) in Escherichia coli (68), Dbh in Sulfolobus acidocaldarius (9, 10), and Dpo4 in Sulfolobus solfataricus (11, 12). A decade of investigations directed at the structural and functional properties of bypass polymerases have significantly improved our understanding of this class of enzymes (5, 13). A unique feature of Y-class polymerases, compared with the common right-handed arrangement of palm, thumb, and finger subdomains of high fidelity (i.e. A-class) DNA polymerases (14), is a “little finger” or “PAD” (palm-associated domain) subdomain that plays a crucial role in lesion bypass (12, 1521). In addition to the little finger subdomain at the C-terminal end of the catalytic core, both Rev1 and Pol κ exhibit an N-terminal extension that is absent in other translesion polymerases. The N-terminal extension in the structure of the ternary (human) hPol κ·DNA·dTTP complex folds into a U-shaped tether-helix-turn-helix “clasp” that is located between the thumb and little finger domains, allowing the polymerase to completely encircle the DNA (18). Although the precise role of the clasp for lesion bypass by hPol κ remains to be established, it is clear that this entity is functionally important, because mutant enzymes with partially or completely removed clasps exhibit diminished catalytic activity compared with the full-length catalytic core (hPol κ N1–526) or a core lacking the N-terminal 19 residues (hPol κ N19–526; the construct used for crystal structure determination of the ternary complex (18)).7,8-Dihydro-8-oxo-2′-deoxyguanosine (8-oxoG), found in both lower organisms and eukaryotes, is a major lesion that is a consequence of oxidative stress. The lesion is of relevance not only because of its association with cancer (22, 23), but also in connection with aging (24), hepatitis (25), and infertility (26). It is far from clear which DNA polymerases bypass 8-oxoG most often in a cellular context, but given the ubiquitous nature of the lesion it seems likely that more than one enzyme could encounter the lesion. Replicative polymerases commonly insert dATP opposite template 8-oxoG, with the lesion adopting the preferred syn conformation (e.g. 27, 28). It was recently found that the translesion polymerase Dpo4 from S. solfataricus synthesizes efficiently past 8-oxoG, inserting ≥95% dCTP > dATP opposite the lesion (29, 30). The efficient and low error bypass of the 8-oxoG lesion by Dpo4 is associated to a large extent with an arginine residue in the little finger domain (17). In the crystal structure of the ternary Dpo4·DNA·dCTP complex, the side chain of Arg-332 forms a hydrogen bond to the 8-oxygen of 8-oxoG, thus shifting the nucleoside conformational equilibrium toward the anti state and enabling a Watson-Crick binding mode with the incoming dCTP (30). The efficient and accurate replication of templates bearing 8-oxoG by yeast Pol η (31, 32) may indicate similarities between the bypass reactions catalyzed by the archaeal and eukaryotic enzymes. In contrast, bypass synthesis opposite 8-oxoG by human Pol κ is error-prone, resulting in efficient incorporation of A (3335). The inaccurate bypass of 8-oxoG is thought to contribute to the deleterious effects associated with the lesion. These observations indicate different behaviors of the eukaryotic trans-lesion Pol κ and its archaeal “homolog” Dpo4 vis-à-vis the major oxidative stress lesion 8-oxoG. A mechanistic understanding of human DNA polymerases that bypass 8-oxoG in an error-prone fashion, such as hPol κ, is therefore of great interest.To elucidate commonalities and differences between the trans-8-oxoG syntheses of S. solfataricus Dpo4, yeast Pol η, and hPol κ, we carried out a comprehensive analysis of the bypass activity for the latter with template·DNA containing the 8-oxoG lesion, including pre-steady-state and steady-state kinetics of primer extension opposite and beyond 8-oxoG and LC-MS/MS assays of full-length extension products. We determined crystal structures of ternary hPol κ-(8-oxoG)DNA-dGTP and hPol κ-(8-oxoG)DNA-dATP complexes, apparently the first for any complex with adducted DNA for the κ enzyme reported to date. Our work demonstrates clear distinctions between genetically related translesion polymerases and provides insights into the origins of the error-prone reactions opposite 8-oxoG catalyzed by Y-family DNA polymerases.  相似文献   

19.
Excessive accumulation of β-amyloid peptides in the brain is a major cause for the pathogenesis of Alzheimer disease. β-Amyloid is derived from β-amyloid precursor protein (APP) through sequential cleavages by β- and γ-secretases, whose enzymatic activities are tightly controlled by subcellular localization. Delineation of how intracellular trafficking of these secretases and APP is regulated is important for understanding Alzheimer disease pathogenesis. Although APP trafficking is regulated by multiple factors including presenilin 1 (PS1), a major component of the γ-secretase complex, and phospholipase D1 (PLD1), a phospholipid-modifying enzyme, regulation of intracellular trafficking of PS1/γ-secretase and β-secretase is less clear. Here we demonstrate that APP can reciprocally regulate PS1 trafficking; APP deficiency results in faster transport of PS1 from the trans-Golgi network to the cell surface and increased steady state levels of PS1 at the cell surface, which can be reversed by restoring APP levels. Restoration of APP in APP-deficient cells also reduces steady state levels of other γ-secretase components (nicastrin, APH-1, and PEN-2) and the cleavage of Notch by PS1/γ-secretase that is more highly correlated with cell surface levels of PS1 than with APP overexpression levels, supporting the notion that Notch is mainly cleaved at the cell surface. In contrast, intracellular trafficking of β-secretase (BACE1) is not regulated by APP. Moreover, we find that PLD1 also regulates PS1 trafficking and that PLD1 overexpression promotes cell surface accumulation of PS1 in an APP-independent manner. Our results clearly elucidate a physiological function of APP in regulating protein trafficking and suggest that intracellular trafficking of PS1/γ-secretase is regulated by multiple factors, including APP and PLD1.An important pathological hallmark of Alzheimer disease (AD)4 is the formation of senile plaques in the brains of patients. The major components of those plaques are β-amyloid peptides (Aβ), whose accumulation triggers a cascade of neurodegenerative steps ending in formation of senile plaques and intraneuronal fibrillary tangles with subsequent neuronal loss in susceptible brain regions (1, 2). Aβ is proteolytically derived from the β-amyloid precursor protein (APP) through sequential cleavages by β-secretase (BACE1), a novel membrane-bound aspartyl protease (3, 4), and by γ-secretase, a high molecular weight complex consisting of at least four components: presenilin (PS), nicastrin (NCT), anterior pharynx-defective-1 (APH-1), and presenilin enhancer-2 (PEN-2) (5, 6). APP is a type I transmembrane protein belonging to a protein family that includes APP-like protein 1 (APLP1) and 2 (APLP2) in mammals (7, 8). Full-length APP is synthesized in the endoplasmic reticulum (ER) and transported through the Golgi apparatus. Most secreted Aβ peptides are generated within the trans-Golgi network (TGN), also the major site of steady state APP in neurons (911). APP can be transported to the cell surface in TGN-derived secretory vesicles if not proteolyzed to Aβ or an intermediate metabolite. At the cell surface APP is either cleaved by α-secretase to produce soluble sAPPα (12) or reinternalized for endosomal/lysosomal degradation (13, 14). Aβ may also be generated in endosomal/lysosomal compartments (15, 16). In contrast to neurotoxic Aβ peptides, sAPPα possesses neuroprotective potential (17, 18). Thus, the subcellular distribution of APP and proteases that process it directly affect the ratio of sAPPα to Aβ, making delineation of the mechanisms responsible for regulating trafficking of all of these proteins relevant to AD pathogenesis.Presenilin (PS) is a critical component of the γ-secretase. Of the two mammalian PS gene homologues, PS1 and PS2, PS1 encodes the major form (PS1) in active γ-secretase (19, 20). Nascent PSs undergo endoproteolytic cleavage to generate an amino-terminal fragment (NTF) and a carboxyl-terminal fragment (CTF) to form a functional PS heterodimer (21). Based on observations that PSs possess two highly conserved aspartate residues indispensable for γ-secretase activity and that specific transition state analogue γ-secretase inhibitors bind to PS1 NTF/CTF heterodimers (5, 22), PSs are believed to be the catalytic component of the γ-secretase complex. PS assembles with three other components, NCT, APH-1, and PEN-2, to form the functional γ-secretase (5, 6). Strong evidence suggests that PS1/γ-secretase resides principally in the ER, early Golgi, TGN, endocytic and intermediate compartments, most of which (except the TGN) are not major subcellular sites for APP (23, 24). In addition to generating Aβ and cleaving APP to release the APP intracellular domain, PS1/γ-secretase cleaves other substrates such as Notch (25), cadherin (26), ErbB4 (27), and CD44 (28), releasing their respective intracellular domains. Interestingly, PS1/γ-secretase cleavage of different substrates seems to occur at different subcellular compartments; APP is mainly cleaved at the TGN and early endosome domains, whereas Notch is predominantly cleaved at the cell surface (9, 11, 29). Thus, perturbing intracellular trafficking of PS1/γ-secretase may alter interactions between PS1/γ-secretase and APP, contributing to either abnormal Aβ generation and AD pathogenesis or decreased access of PS1/γ-secretase to APP such that Aβ production is reduced. However, mechanisms regulating PS1/γ-secretase trafficking warrant further investigation.In addition to participating in γ-secretase activity, PS1 regulates intracellular trafficking of several membrane proteins, including other γ-secretase components (nicastrin, APH-1, and PEN-2) and the substrate APP (reviewed in Ref. 30). Intracellular APP trafficking is highly regulated and requires other factors such as mint family members and SorLA (2). Moreover, we recently found that phospholipase D1 (PLD1), a phospholipid-modifying enzyme that regulates membrane trafficking events, can interact with PS1, and can regulate budding of APP-containing vesicles from the TGN and delivery of APP to the cell surface (31, 32). Interestingly, Kamal et al. (33) identified an axonal membrane compartment that contains APP, BACE1, and PS1 and showed that fast anterograde axonal transport of this compartment is mediated by APP and kinesin-I, implying a traffic-regulating role for APP. Increased APP expression is also shown to decrease retrograde axonal transport of nerve growth factor (34). However, whether APP indeed regulates intracellular trafficking of proteins including BACE1 and PS1/γ-secretase requires further validation. In the present study we demonstrate that intracellular trafficking of PS1, as well as that of other γ-secretase components, but not BACE1, is regulated by APP. APP deficiency promotes cell surface delivery of PS1/γ-secretase complex and facilitates PS1/γ-secretase-mediated Notch cleavage. In addition, we find that PLD1 also regulates intracellular trafficking of PS1 through a different mechanism and more potently than APP.  相似文献   

20.
Poly(ADP-ribose) polymerase-1 (PARP-1), a eucaryotic nuclear DNA-binding protein that is activated by breaks in DNA chains, may be involved in the base excision repair (BER) because DNAs containing single-stranded gaps and breaks are intermediates of BER. The effect of PARP-1 on the DNA synthesis catalyzed in vitro by DNA polymerase beta (pol beta) was studied using analogs of DNA substrates produced during BER and imitating intermediates of the short patch and long patch subpathways of BER. Oligonucleotide duplexes of 34 bp that contained a mononucleotide gap or a single-strand break with tetrahydrofuran phosphate or phosphate at the 5;-end of the downstream oligonucleotide were taken as DNA substrates. The efficiency of DNA synthesis was determined at various ratios of pol beta and PARP-1. The efficiency of gap filling was decreased in the presence of PARP-1, but strand-displacement DNA synthesis was inhibited significantly stronger, which seemed to be due to competition between PARP-1 and pol beta for DNA. In the presence of NAD+ and single-strand breaks in DNA, PARP-1 catalyzes the synthesis of poly(ADP-ribose) covalently attached to the enzyme, and this automodification is thought to provide for dissociation of PARP-1 from DNA. The effect of PARP-1 automodification on inhibition of DNA synthesis was studied, and efficiency of mononucleotide gap filling was shown to be restored, but strand-displacement synthesis did not revert to the level observed in the absence of PARP-1. PARP-1 is suggested to regulate the interaction between pol beta and DNA, in particular, via its own automodification.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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