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 593 毫秒
1.
We report here a biochemical and structural characterization of domain 2 of the nonstructural 5A protein (NS5A) from the JFH1 Hepatitis C virus strain and its interactions with cyclophilins A and B (CypA and CypB). Gel filtration chromatography, circular dichroism spectroscopy, and finally NMR spectroscopy all indicate the natively unfolded nature of this NS5A-D2 domain. Because mutations in this domain have been linked to cyclosporin A resistance, we used NMR spectroscopy to investigate potential interactions between NS5A-D2 and cellular CypA and CypB. We observed a direct molecular interaction between NS5A-D2 and both cyclophilins. The interaction surface on the cyclophilins corresponds to their active site, whereas on NS5A-D2, it proved to be distributed over the many proline residues of the domain. NMR heteronuclear exchange spectroscopy yielded direct evidence that many proline residues in NS5A-D2 form a valid substrate for the enzymatic peptidyl-prolyl cis/trans isomerase (PPIase) activity of CypA and CypB.Hepatitis C virus (HCV)4 is a small, positive strand, RNA-enveloped virus belonging to the Flaviviridae family and the genus Hepacivirus. With 120–180 million chronically infected individuals worldwide, hepatitis C virus infection represents a major cause of chronic hepatitis, liver cirrhosis, and hepatocellular carcinoma (1). The HCV viral genome (∼9.6 kb) codes for a unique polyprotein of ∼3000 amino acids (recently reviewed in Refs. 24). Following processing via viral and cellular proteases, this polyprotein gives rise to at least 10 viral proteins, divided into structural (core, E1, and E2 envelope glycoproteins) and nonstructural proteins (p7, NS2, NS3, NS4A, NS4B, NS5A, NS5B). Nonstructural proteins are involved in polyprotein processing and viral replication. The set composed of NS3, NS4A, NS4B, NS5A, and NS5B constitutes the minimal protein component required for viral replication (5).Cyclophilins are cellular proteins that have been identified first as CsA-binding proteins (6). As FK506-binding proteins (FKBP) and parvulins, cyclophilins are peptidyl-prolyl cis/trans isomerases (PPIase) that catalyze the cis/trans isomerization of the peptide linkage preceding a proline (6, 7). Several subtypes of cyclophilins are present in mammalian cells (8). They share a high sequence homology and a well conserved three-dimensional structure but display significant differences in their primary cellular localization and in abundance (9). CypA, the most abundant of the cyclophilins, is primarily cytoplasmic, whereas CypB is directed to the endoplasmic reticulum lumen or the secretory pathway. CypD, on the other hand, is the mitochondrial cyclophilin. Cyclophilins are involved in numerous physiological processes such as protein folding, immune response, and apoptosis and also in the replication cycle of viruses including vaccinia virus, vesicular stomatitis virus, severe acute respiratory syndrome (SARS)-coronavirus, and human immunodeficiency virus (HIV) (for review see Ref. 10). For HIV, CypA has been shown to interact with the capsid domain of the HIV Gag precursor polyprotein (11). CypA thereby competes with capsid domain/TRIM5 interaction, resulting in a loss of the antiviral protective effect of the cellular restriction factor TRIM5α (12, 13). Moreover, it has been shown that CypA catalyzes the cis/trans isomerization of Gly221-Pro222 in the capsid domain and that it has functional consequences for HIV replication efficiency (1416). For HCV, Watashi et al. (17) have described a molecular and functional interaction between NS5B, the viral RNA-dependent RNA polymerase (RdRp), and cyclophilin B (CypB). CypB may be a key regulator in HCV replication by modulating the affinity of NS5B for RNA. This regulation is abolished in the presence of cyclosporin A (CsA), an inhibitor of cyclophilins (6). These results provided for the first time a molecular mechanism for the early-on observed anti-HCV activity of CsA (1820). Although this initial report suggests that only CypB would be involved in the HCV replication process (17), a growing number of studies have recently pointed out a role for other cyclophilins (2125).In vitro selection of CsA-resistant HCV mutants indicated the importance of two HCV nonstructural proteins, NS5B and NS5A (26), with a preponderant effect for mutations in the C-terminal half of NS5A. NS5A is a large phosphoprotein (49 kDa), indispensable for HCV replication and particle assembly (2729), but for which the exact function(s) in the HCV replication cycle remain to be elucidated. This nonstructural protein is anchored to the cytoplasmic leaflet of the endoplasmic reticulum membrane via an N-terminal amphipathic α-helix (residues 1–27) (30, 31). Its cytoplasmic sequence can be divided into three domains: D1 (residues 27–213), D2 (residues 250–342), and D3 (residues 356–447), all connected by low complexity sequences (32). D1, a zinc-binding domain, adopts a dimeric claw-shaped structure, which is proposed to interact with RNA (33, 34). NS5A-D2 is essential for HCV replication, whereas NS5A-D3 is a key determinant for virus infectious particle assembly (27, 35). NS5A-D2 and -D3, for which sequence conservation among HCV genotypes is significantly lower than for D1, have been proposed to be natively unfolded domains (28, 32). Molecular and structural characterization of NS5A-D2 from HCV genotype 1a has confirmed the disordered nature of this domain (36, 37).As it is still not clear which cyclophilins are cofactors for HCV replication, and as mutations in HCV NS5A protein have been associated with CsA resistance, we decided to examine the interaction between both CypA and CypB and domain 2 of the HCV NS5A protein. We first characterized, at the molecular level, NS5A-D2 from the HCV JFH1 infectious strain (genotype 2a) and showed by NMR spectroscopy that this natively unfolded domain indeed interacts with both cyclophilin A and cyclophilin B. Our NMR chemical shift mapping experiments indicated that the interaction occurs at the level of the cyclophilin active site, whereas it lacks a precise localization on NS5A-D2. A peptide derived from the only well conserved amino acid motif in NS5A-D2 did interact with cyclophilin A but only with a 10-fold lower affinity than the full domain. We concluded from this that the many proline residues form multiple anchoring points, especially when they adopt the cis conformation. NMR exchange spectroscopy further demonstrated that NS5A-D2 is a substrate for the PPIase activities of both CypA and CypB. Both the NS5A/cyclophilin interaction and the PPIase activity of the cyclophilins on NS5A-D2 were abolished by CsA, underscoring the specificity of the interaction.  相似文献   

2.
Early events leading to the establishment of hepatitis C virus (HCV) infection are not completely understood. We show that intact and dynamic microtubules play a key role in the initiation of productive HCV infection. Microtubules were required for virus entry into cells, as evidenced using virus pseudotypes presenting HCV envelope proteins on their surface. Studies carried out using the recent infectious HCV model revealed that microtubules also play an essential role in early, postfusion steps of the virus cycle. Moreover, low concentrations of vinblastin and nocodazol, microtubule-affecting drugs, and paclitaxel, which stabilizes microtubules, inhibited infection, suggesting that microtubule dynamic instability and/or treadmilling mechanisms are involved in HCV internalization and early transport. By protein chip and direct core-dependent pull-down assays, followed by mass spectrometry, we identified β- and α-tubulin as cellular partners of the HCV core protein. Surface plasmon resonance analyses confirmed that core directly binds to tubulin with high affinity via amino acids 2-117. The interaction of core with tubulin in vitro promoted its polymerization and enhanced the formation of microtubules. Immune electron microscopy showed that HCV core associates, at least temporarily, with microtubules polymerized in its presence. Studies by confocal microscopy showed a juxtaposition of core with microtubules in HCV-infected cells. In summary, we report that intact and dynamic microtubules are required for virus entry into cells and for early postfusion steps of infection. HCV may exploit a direct interaction of core with tubulin, enhancing microtubule polymerization, to establish efficient infection and promote virus transport and/or assembly in infected cells.HCV5 infection is a major cause of chronic liver disease, which frequently progresses to cirrhosis and hepatocellular carcinoma. HCV represents a global public health problem, with 130 million people infected worldwide. There is currently no vaccine directed against HCV and the available antiviral treatments eliminate the virus in 40-80% of patients, depending on the virus genotype (for review, see Ref. 1).HCV has a single-stranded, positive-sense RNA genome of ∼9.6 kilobases encoding a large polyprotein that is processed by both host and viral proteases to produce three structural proteins (core protein and the envelope glycoproteins E1 and E2), p7, and six nonstructural proteins, which are involved in polyprotein processing and replication of the virus genome (for review, see Ref. 2).HCV core is a basic protein, synthesized as the most N-terminal component of the polyprotein, and is followed by the signal sequence of the E1 envelope glycoprotein (3). The polypeptide is cleaved by signal peptidase and signal peptide peptidase, resulting in the release of core from the endoplasmic reticulum membrane and its trafficking to lipid droplets (3-5). Mature core protein forms the viral nucleocapsid (6) and consists of two domains, D1 and D2. D1 lies at the protein N terminus, is composed of about 117 amino acids (aa), and is involved in RNA binding (7). D2 is relatively hydrophobic, has a length of about 55 aa, and targets HCV core to lipid droplets (8).Microtubules (MTs) are ubiquitous cytoskeleton components that play a key role in various cellular processes relating to cell shape and division, motility, and intracellular trafficking (9). MTs are dynamic, polarized polymers composed of α/β-tubulin heterodimers that undergo alternate phases of growth and shrinkage, dependent on so-called “dynamic instability” (10). Active transport by MTs is bidirectional and involves both plus and minus end-directed motors: kinesin and dynein (11, 12).Another mechanism of cytosolic transport on MTs, called “treadmilling” (13, 14) involves polymerization at the plus end and depolymerization at the minus end after severing of MTs by cellular katenin (15).MTs have important functions in the life cycle of most viruses (13, 16, 17). Cytoplasmic transport on MTs provides viruses with the means to reach sites of replication or enables progeny virus to leave the infected cell. Some viruses, such as Ebola virus (18) or reovirus (19), are transported on MTs within membranous compartments, whereas other viruses like herpes simplex virus type 1 (20), murine polyoma virus (21), human cytomegalovirus (22), or adenovirus (23) interact with MT motors or MT-associated proteins to allow their transport along microtubules.Previous studies have established that the cell cytoskeleton is involved in HCV replication, since HCV replication complexes are subjected to intracellular transport and their formation is closely linked to the dynamic organization of endoplasmic reticulum, actin filaments, and the microtubule network (24-26). In addition, intact microtubules are essential for viral morphogenesis and the secretion of progeny virus from infected cells (27). The role of microtubules in HCV cell entry and the initiation of productive HCV infection has not yet been addressed.In this study, we provide evidence that the MT network plays a key role in HCV cell entry and postfusion steps of the virus cycle that lead to the establishment of productive HCV infection. The initial steps of the viral cycle are sensitive to MT-affecting drugs that inhibit MT formation or depolymerize or stabilize microtubules. We also show a unique property of the HCV core protein, its capacity to directly bind to tubulin and to enhance MT polymerization in vitro. Our findings suggest that HCV could exploit the MT network by polymerization-related mechanisms to productively infect its target cell. Thus, microtubules may provide a novel target for therapeutic interventions against HCV infection.  相似文献   

3.
4.
5.
6.
7.
As obligate intracellular parasites, viruses exploit diverse cellular signaling machineries, including the mitogen-activated protein-kinase pathway, during their infections. We have demonstrated previously that the open reading frame 45 (ORF45) of Kaposi sarcoma-associated herpesvirus interacts with p90 ribosomal S6 kinases (RSKs) and strongly stimulates their kinase activities (Kuang, E., Tang, Q., Maul, G. G., and Zhu, F. (2008) J. Virol. 82 ,1838 -1850). Here, we define the mechanism by which ORF45 activates RSKs. We demonstrated that binding of ORF45 to RSK increases the association of extracellular signal-regulated kinase (ERK) with RSK, such that ORF45, RSK, and ERK formed high molecular mass protein complexes. We further demonstrated that the complexes shielded active pERK and pRSK from dephosphorylation. As a result, the complex-associated RSK and ERK were activated and sustained at high levels. Finally, we provide evidence that this mechanism contributes to the sustained activation of ERK and RSK in Kaposi sarcoma-associated herpesvirus lytic replication.The extracellular signal-regulated kinase (ERK)2 mitogen-activated protein kinase (MAPK) signaling pathway has been implicated in diverse cellular physiological processes including proliferation, survival, growth, differentiation, and motility (1-4) and is also exploited by a variety of viruses such as Kaposi sarcoma-associated herpesvirus (KSHV), human cytomegalovirus, human immunodeficiency virus, respiratory syncytial virus, hepatitis B virus, coxsackie, vaccinia, coronavirus, and influenza virus (5-17). The MAPK kinases relay the extracellular signaling through sequential phosphorylation to an array of cytoplasmic and nuclear substrates to elicit specific responses (1, 2, 18). Phosphorylation of MAPK is reversible. The kinetics of deactivation or duration of signaling dictates diverse biological outcomes (19, 20). For example, sustained but not transient activation of ERK signaling induces the differentiation of PC12 cells into sympathetic-like neurons and transformation of NIH3T3 cells (20-22). During viral infection, a unique biphasic ERK activation has been observed for some viruses (an early transient activation triggered by viral binding or entry and a late sustained activation correlated with viral gene expression), but the responsible viral factors and underlying mechanism for the sustained ERK activation remain largely unknown (5, 8, 13, 23).The p90 ribosomal S6 kinases (RSKs) are a family of serine/threonine kinases that lie at the terminus of the ERK pathway (1, 24-26). In mammals, four isoforms are known, RSK1 to RSK4. Each one has two catalytically functional kinase domains, the N-terminal kinase domain (NTKD) and C-terminal kinase domain (CTKD) as well as a linker region between the two. The NTKD is responsible for phosphorylation of exogenous substrates, and the CTKD and linker region regulate RSK activation (1, 24, 25). In quiescent cells ERK binds to the docking site in the C terminus of RSK (27-29). Upon mitogen stimulation, ERK is activated by its upstream MAPK/ERK kinase (MEK). The active ERK phosphorylates Thr-359/Ser-363 of RSK in the linker region (amino acid numbers refer to human RSK1) and Thr-573 in the CTKD activation loop. The activated CTKD then phosphorylates Ser-380 in the linker region, creating a docking site for 3-phosphoinositide-dependent protein kinase-1. The 3-phosphoinositide-dependent protein kinase-1 phosphorylates Ser-221 of RSK in the activation loop and activates the NTKD. The activated NTKD autophosphorylates the serine residue near the ERK docking site, causing a transient dissociation of active ERK from RSK (25, 26, 28). The stimulation of quiescent cells by a mitogen such as epidermal growth factor or a phorbol ester such as 12-O-tetradecanoylphorbol-13-acetate (TPA) usually results in a transient RSK activation that lasts less than 30 min. RSKs have been implicated in regulating cell survival, growth, and proliferation. Mutation or aberrant expression of RSK has been implicated in several human diseases including Coffin-Lowry syndrome and prostate and breast cancers (1, 24, 25, 30-32).KSHV is a human DNA tumor virus etiologically linked to Kaposi sarcoma, primary effusion lymphoma, and a subset of multicentric Castleman disease (33, 34). Infection and reactivation of KSHV activate multiple MAPK pathways (6, 12, 35). Noticeably, the ERK/RSK activation is sustained late during KSHV primary infection and reactivation from latency (5, 6, 12, 23), but the mechanism of the sustained ERK/RSK activation is unclear. Recently, we demonstrated that ORF45, an immediate early and also virion tegument protein of KSHV, interacts with RSK1 and RSK2 and strongly stimulates their kinase activities (23). We also demonstrated that the activation of RSK plays an essential role in KSHV lytic replication (23). In the present study we determined the mechanism of ORF45-induced sustained ERK/RSK activation. We found that ORF45 increases the association of RSK with ERK and protects them from dephosphorylation, causing sustained activation of both ERK and RSK.  相似文献   

8.
For Bovine viral diarrhea virus (BVDV), the type species of the genus Pestivirus in the family Flaviviridae, cytopathogenic (cp) and noncytopathogenic (ncp) viruses are distinguished according to their effect on cultured cells. It has been established that cytopathogenicity of BVDV correlates with efficient production of viral nonstructural protein NS3 and with enhanced viral RNA synthesis. Here, we describe generation and characterization of a temperature-sensitive (ts) mutant of cp BVDV strain CP7, termed TS2.7. Infection of bovine cells with TS2.7 and the parent CP7 at 33°C resulted in efficient viral replication and a cytopathic effect. In contrast, the ability of TS2.7 to cause cytopathogenicity at 39.5°C was drastically reduced despite production of high titers of infectious virus. Further experiments, including nucleotide sequencing of the TS2.7 genome and reverse genetics, showed that a Y1338H substitution at residue 193 of NS2 resulted in the temperature-dependent attenuation of cytopathogenicity despite high levels of infectious virus production. Interestingly, TS2.7 and the reconstructed mutant CP7-Y1338H produced NS3 in addition to NS2-3 throughout infection. Compared to the parent CP7, NS2-3 processing was slightly decreased at both temperatures. Quantification of viral RNAs that were accumulated at 10 h postinfection demonstrated that attenuation of the cytopathogenicity of the ts mutants at 39.5°C correlated with reduced amounts of viral RNA, while the efficiency of viral RNA synthesis at 33°C was not affected. Taken together, the results of this study show that a mutation in BVDV NS2 attenuates viral RNA replication and suppresses viral cytopathogenicity at high temperature without altering NS3 expression and infectious virus production in a temperature-dependent manner.The pestiviruses Bovine viral diarrhea virus-1 (BVDV-1), BVDV-2, Classical swine fever virus (CSFV), and Border disease virus (BDV) are causative agents of economically important livestock diseases. Together with the genera Flavivirus, including several important human pathogens like Dengue fever virus, West Nile virus, Yellow fever virus, and Tick-borne encephalitis virus, and Hepacivirus (human Hepatitis C virus [HCV]), the genus Pestivirus constitutes the family Flaviviridae (8, 20). All members of this family are enveloped viruses with a single-stranded positive-sense RNA genome encompassing one large open reading frame (ORF) flanked by 5′ and 3′ nontranslated regions (NTR) (see references 8 and 28 for reviews). The ORF encodes a polyprotein which is co- and posttranslationally processed into the mature viral proteins by viral and cellular proteases. For BVDV, the RNA genome is about 12.3 kb in length and encodes a polyprotein of about 3,900 amino acids. The first third of the ORF encodes a nonstructural (NS) autoprotease and four structural proteins, while the remaining part of the genome encodes NS proteins which share many common characteristics and functions with the corresponding NS proteins encoded by the HCV genome (8, 28). NS2 of BVDV represents a cysteine autoprotease which is distantly related to the HCV NS2-3 protease (26). NS3, NS4A, NS4B, NS5A, and NS5B are essential components of the pestivirus replicase (7, 10, 49). NS3 possesses multiple enzymatic activities, namely serine protease (48, 52, 53), NTPase (46), and helicase activity (51). NS4A acts as an essential cofactor for the NS3 proteinase. NS5B represents the RNA-dependent RNA polymerase (RdRp) (22, 56). The functions of NS4B and NS5A remain to be determined. NS5A has been shown to be a phosphorylated protein that is associated with cellular serine/threonine kinases (44).According to their effects in tissue culture, two biotypes of pestiviruses are distinguished: cytopathogenic (cp) and noncytopathogenic (ncp) viruses (17, 27). The occurrence of cp BVDV in cattle persistently infected with ncp BVDV is directly linked to the induction of lethal mucosal disease in cattle (12, 13). Previous studies have shown that cp BVDV strains evolved from ncp BVDV strains by different kinds of mutations. These include RNA recombination with various cellular mRNAs, resulting in insertions of cellular protein-coding sequences into the viral genome, as well as insertions, duplications, and deletions of viral sequences, and point mutations (1, 2, 9, 24, 33, 36, 37, 42). A common consequence of all these genetic changes in cp BVDV genomes is the efficient production of NS3 at early and late phases of infection. In contrast, NS3 cannot be detected in cells at late time points after infection with ncp BVDV. An additional major difference is that the cp viruses produce amounts of viral RNA significantly larger than those of their ncp counterparts (7, 32, 50). While there is clear evidence that cell death induced by cp BVDV is mediated by apoptosis, the molecular mechanisms involved in pestiviral cytopathogenicity are poorly understood. In particular, the role of NS3 in triggering apoptosis remains unclear. It has been hypothesized that the NS3 serine proteinase might be involved in activation of the apoptotic proteolytic cascade (21, 55). Furthermore, it has been suggested that the NS3-mediated, enhanced viral RNA synthesis of cp BVDV and subsequently larger amounts of viral double-stranded RNAs may play a crucial role in triggering apoptosis (31, 54).In this study, we describe generation and characterization of a temperature-sensitive (ts) cp BVDV mutant whose ability to cause viral cytopathogenicity at high temperature is strongly attenuated. Our results demonstrate that a single amino acid substitution in NS2 attenuates BVDV cytopathogenicity at high temperature without affecting production of infectious viruses and expression of NS3 in a temperature-dependent manner.  相似文献   

9.
10.
Mitochondrial dysregulation is strongly implicated in Parkinson disease. Mutations in PTEN-induced kinase 1 (PINK1) are associated with familial parkinsonism and neuropsychiatric disorders. Although overexpressed PINK1 is neuroprotective, less is known about neuronal responses to loss of PINK1 function. We found that stable knockdown of PINK1 induced mitochondrial fragmentation and autophagy in SH-SY5Y cells, which was reversed by the reintroduction of an RNA interference (RNAi)-resistant plasmid for PINK1. Moreover, stable or transient overexpression of wild-type PINK1 increased mitochondrial interconnectivity and suppressed toxin-induced autophagy/mitophagy. Mitochondrial oxidant production played an essential role in triggering mitochondrial fragmentation and autophagy in PINK1 shRNA lines. Autophagy/mitophagy served a protective role in limiting cell death, and overexpressing Parkin further enhanced this protective mitophagic response. The dominant negative Drp1 mutant inhibited both fission and mitophagy in PINK1-deficient cells. Interestingly, RNAi knockdown of autophagy proteins Atg7 and LC3/Atg8 also decreased mitochondrial fragmentation without affecting oxidative stress, suggesting active involvement of autophagy in morphologic remodeling of mitochondria for clearance. To summarize, loss of PINK1 function elicits oxidative stress and mitochondrial turnover coordinated by the autophagic and fission/fusion machineries. Furthermore, PINK1 and Parkin may cooperate through different mechanisms to maintain mitochondrial homeostasis.Parkinson disease is an age-related neurodegenerative disease that affects ∼1% of the population worldwide. The causes of sporadic cases are unknown, although mitochondrial or oxidative toxins such as 1-methyl-4-phenylpyridinium, 6-hydroxydopamine (6-OHDA),3 and rotenone reproduce features of the disease in animal and cell culture models (1). Abnormalities in mitochondrial respiration and increased oxidative stress are observed in cells and tissues from parkinsonian patients (2, 3), which also exhibit increased mitochondrial autophagy (4). Furthermore, mutations in parkinsonian genes affect oxidative stress response pathways and mitochondrial homeostasis (5). Thus, disruption of mitochondrial homeostasis represents a major factor implicated in the pathogenesis of sporadic and inherited parkinsonian disorders (PD).The PARK6 locus involved in autosomal recessive and early-onset PD encodes for PTEN-induced kinase 1 (PINK1) (6, 7). PINK1 is a cytosolic and mitochondrially localized 581-amino acid serine/threonine kinase that possesses an N-terminal mitochondrial targeting sequence (6, 8). The primary sequence also includes a putative transmembrane domain important for orientation of the PINK1 domain (8), a conserved kinase domain homologous to calcium calmodulin kinases, and a C-terminal domain that regulates autophosphorylation activity (9, 10). Overexpression of wild-type PINK1, but not its PD-associated mutants, protects against several toxic insults in neuronal cells (6, 11, 12). Mitochondrial targeting is necessary for some (13) but not all of the neuroprotective effects of PINK1 (14), implicating involvement of cytoplasmic targets that modulate mitochondrial pathobiology (8). PINK1 catalytic activity is necessary for its neuroprotective role, because a kinase-deficient K219M substitution in the ATP binding pocket of PINK1 abrogates its ability to protect neurons (14). Although PINK1 mutations do not seem to impair mitochondrial targeting, PD-associated mutations differentially destabilize the protein, resulting in loss of neuroprotective activities (13, 15).Recent studies indicate that PINK1 and Parkin interact genetically (3, 16-18) to prevent oxidative stress (19, 20) and regulate mitochondrial morphology (21). Primary cells derived from PINK1 mutant patients exhibit mitochondrial fragmentation with disorganized cristae, recapitulated by RNA interference studies in HeLa cells (3).Mitochondria are degraded by macroautophagy, a process involving sequestration of cytoplasmic cargo into membranous autophagic vacuoles (AVs) for delivery to lysosomes (22, 23). Interestingly, mitochondrial fission accompanies autophagic neurodegeneration elicited by the PD neurotoxin 6-OHDA (24, 25). Moreover, mitochondrial fragmentation and increased autophagy are observed in neurodegenerative diseases including Alzheimer and Parkinson diseases (4, 26-28). Although inclusion of mitochondria in autophagosomes was once believed to be a random process, as observed during starvation, studies involving hypoxia, mitochondrial damage, apoptotic stimuli, or limiting amounts of aerobic substrates in facultative anaerobes support the concept of selective mitochondrial autophagy (mitophagy) (29, 30). In particular, mitochondrially localized kinases may play an important role in models involving oxidative mitochondrial injury (25, 31, 32).Autophagy is involved in the clearance of protein aggregates (33-35) and normal regulation of axonal-synaptic morphology (36). Chronic disruption of lysosomal function results in accumulation of subtly impaired mitochondria with decreased calcium buffering capacity (37), implicating an important role for autophagy in mitochondrial homeostasis (37, 38). Recently, Parkin, which complements the effects of PINK1 deficiency on mitochondrial morphology (3), was found to promote autophagy of depolarized mitochondria (39). Conversely, Beclin 1-independent autophagy/mitophagy contributes to cell death elicited by the PD toxins 1-methyl-4-phenylpyridinium and 6-OHDA (25, 28, 31, 32), causing neurite retraction in cells expressing a PD-linked mutation in leucine-rich repeat kinase 2 (40). Whereas properly regulated autophagy plays a homeostatic and neuroprotective role, excessive or incomplete autophagy creates a condition of “autophagic stress” that can contribute to neurodegeneration (28).As mitochondrial fragmentation (3) and increased mitochondrial autophagy (4) have been described in human cells or tissues of PD patients, we investigated whether or not the engineered loss of PINK1 function could recapitulate these observations in human neuronal cells (SH-SY5Y). Stable knockdown of endogenous PINK1 gave rise to mitochondrial fragmentation and increased autophagy and mitophagy, whereas stable or transient overexpression of PINK1 had the opposite effect. Autophagy/mitophagy was dependent upon increased mitochondrial oxidant production and activation of fission. The data indicate that PINK1 is important for the maintenance of mitochondrial networks, suggesting that coordinated regulation of mitochondrial dynamics and autophagy limits cell death associated with loss of PINK1 function.  相似文献   

11.
12.
Hepatocellular carcinoma (HCC) is one of the most common and aggressive human malignancies. Recombinant tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) is a promising anti-tumor agent. However, many HCC cells show resistance to TRAIL-induced apoptosis. In this study, we showed that bortezomib, a proteasome inhibitor, overcame TRAIL resistance in HCC cells, including Huh-7, Hep3B, and Sk-Hep1. The combination of bortezomib and TRAIL restored the sensitivity of HCC cells to TRAIL-induced apoptosis. Comparing the molecular change in HCC cells treated with these agents, we found that down-regulation of phospho-Akt (P-Akt) played a key role in mediating TRAIL sensitization of bortezomib. The first evidence was that bortezomib down-regulated P-Akt in a dose- and time-dependent manner in TRAIL-treated HCC cells. Second, LY294002, a PI3K inhibitor, also sensitized resistant HCC cells to TRAIL-induced apoptosis. Third, knocking down Akt1 by small interference RNA also enhanced TRAIL-induced apoptosis in Huh-7 cells. Finally, ectopic expression of mutant Akt (constitutive active) in HCC cells abolished TRAIL sensitization effect of bortezomib. Moreover, okadaic acid, a protein phosphatase 2A (PP2A) inhibitor, reversed down-regulation of P-Akt in bortezomib-treated cells, and PP2A knockdown by small interference RNA also reduced apoptosis induced by the combination of TRAIL and bortezomib, indicating that PP2A may be important in mediating the effect of bortezomib on TRAIL sensitization. Together, bortezomib overcame TRAIL resistance at clinically achievable concentrations in hepatocellular carcinoma cells, and this effect is mediated at least partly via inhibition of the PI3K/Akt pathway.Hepatocellular carcinoma (HCC)2 is currently the fifth most common solid tumor worldwide and the fourth leading cause of cancer-related death. To date, surgery is still the only curative treatment but is only feasible in a small portion of patients (1). Drug treatment is the major therapy for patients with advanced stage disease. Unfortunately, the response rate to traditional chemotherapy for HCC patients is unsatisfactory (1). Novel pharmacological therapy is urgently needed for patients with advanced HCC. In this regard, the approval of sorafenib might open a new era of molecularly targeted therapy in the treatment of HCC patients.Tumor necrosis factor (TNF)-related apoptosis-inducing ligand (TRAIL), a type II transmembrane protein and a member of the TNF family, is a promising anti-tumor agent under clinical investigation (2). TRAIL functions by engaging its receptors expressed on the surface of target cells. Five receptors specific for TRAIL have been identified, including DR4/TRAIL-R1, DR5/TRAIL-R2, DcR1, DcR2, and osteoprotegerin. Among TRAIL receptors, only DR4 and DR5 contain an effective death domain that is essential to formation of death-inducing signaling complex (DISC), a critical step for TRAIL-induced apoptosis. Notably, the trimerization of the death domains recruits an adaptor molecule, Fas-associated protein with death domain (FADD), which subsequently recruits and activates caspase-8. In type I cells, activation of caspase-8 is sufficient to activate caspase-3 to induce apoptosis; however, in another type of cells (type II), the intrinsic mitochondrial pathway is essential for apoptosis characterized by cleavage of Bid and release of cytochrome c from mitochondria, which subsequently activates caspase-9 and caspase-3 (3).Although TRAIL induces apoptosis in malignant cells but sparing normal cells, some tumor cells are resistant to TRAIL-induced apoptosis. Mechanisms responsible for the resistance include receptors and intracellular resistance. Although the cell surface expression of DR4 or DR5 is absolutely required for TRAIL-induced apoptosis, tumor cells expressing these death receptors are not always sensitive to TRAIL due to intracellular mechanisms. For example, the cellular FLICE-inhibitory protein (c-FLIP), a homologue to caspase-8 but without protease activity, has been linked to TRAIL resistance in several studies (4, 5). In addition, inactivation of Bax, a proapoptotic Bcl-2 family protein, resulted in resistance to TRAIL in MMR-deficient tumors (6, 7), and reintroduction of Bax into Bax-deficient cells restored TRAIL sensitivity (8), indicating that the Bcl-2 family plays a critical role in intracellular mechanisms for resistance of TRAIL.Bortezomib, a proteasome inhibitor approved clinically for multiple myeloma and mantle cell lymphoma, has been investigated intensively for many types of cancer (9). Accumulating studies indicate that the combination of bortezomib and TRAIL overcomes the resistance to TRAIL in various types of cancer, including acute myeloid leukemia (4), lymphoma (1013), prostate (1417), colon (15, 18, 19), bladder (14, 16), renal cell carcinoma (20), thyroid (21), ovary (22), non-small cell lung (23, 24), sarcoma (25), and HCC (26, 27). Molecular targets responsible for the sensitizing effect of bortezomib on TRAIL-induced cell death include DR4 (14, 27), DR5 (14, 20, 2223, 28), c-FLIP (4, 11, 2123, 29), NF-κB (12, 24, 30), p21 (16, 21, 25), and p27 (25). In addition, Bcl-2 family also plays a role in the combinational effect of bortezomib and TRAIL, including Bcl-2 (10, 21), Bax (13, 22), Bak (27), Bcl-xL (21), Bik (18), and Bim (15).Recently, we have reported that Akt signaling is a major molecular determinant in bortezomib-induced apoptosis in HCC cells (31). In this study, we demonstrated that bortezomib overcame TRAIL resistance in HCC cells through inhibition of the PI3K/Akt pathway.  相似文献   

13.
Chromosomal abnormalities are frequently caused by problems encountered during DNA replication. Although the ATR-Chk1 pathway has previously been implicated in preventing the collapse of stalled replication forks into double-strand breaks (DSB), the importance of the response to fork collapse in ATR-deficient cells has not been well characterized. Herein, we demonstrate that, upon stalled replication, ATR deficiency leads to the phosphorylation of H2AX by ATM and DNA-PKcs and to the focal accumulation of Rad51, a marker of homologous recombination and fork restart. Because H2AX has been shown to play a facilitative role in homologous recombination, we hypothesized that H2AX participates in Rad51-mediated suppression of DSBs generated in the absence of ATR. Consistent with this model, increased Rad51 focal accumulation in ATR-deficient cells is largely dependent on H2AX, and dual deficiencies in ATR and H2AX lead to synergistic increases in chromatid breaks and translocations. Importantly, the ATM and DNA-PK phosphorylation site on H2AX (Ser139) is required for genome stabilization in the absence of ATR; therefore, phosphorylation of H2AX by ATM and DNA-PKcs plays a pivotal role in suppressing DSBs during DNA synthesis in instances of ATR pathway failure. These results imply that ATR-dependent fork stabilization and H2AX/ATM/DNA-PKcs-dependent restart pathways cooperatively suppress double-strand breaks as a layered response network when replication stalls.Genome maintenance prevents mutations that lead to cancer and age-related diseases. A major challenge in preserving genome integrity occurs in the simple act of DNA replication, in which failures at numerous levels can occur. Besides the mis-incorporation of nucleotides, it is during this phase of the cell cycle that the relatively stable double-stranded nature of DNA is temporarily suspended at the replication fork, a structure that is susceptible to collapse into DSBs.2 Replication fork stability is maintained by a variety of mechanisms, including activation of the ATR-dependent checkpoint pathway.The ATR pathway is activated upon the generation and recognition of extended stretches of single-stranded DNA at stalled replication forks (1-4). Genome maintenance functions for ATR and orthologs in yeast were first indicated by increased chromatid breaks in ATR-/- cultured cells (5) and by the “cut” phenotype observed in Mec1 (Saccharomyces cerevisiae) and Rad3 (Schizosaccharomyces pombe) mutants (6-9). Importantly, subsequent studies in S. cerevisiae demonstrated that mutation of Mec1 or the downstream checkpoint kinase Rad53 led to increased chromosome breaks at regions of the genome that are inherently difficult to replicate (10), and a decreased ability to reinitiate replication fork progression following DNA damage or deoxyribonucleotide depletion (11-14).In vertebrates, similar replication fork stabilizing functions have been demonstrated for ATR and the downstream protein kinase Chk1 (15-20). Several possible mechanisms have been put forward to explain how ATR-Chk1 and orthologous pathways in yeast maintain replication fork stability, including maintenance of replicative polymerases (α, δ, and ε) at forks (17, 21), regulation of branch migrating helicases, such as Blm (22-25), and regulation of homologous recombination, either positively or negatively (26-29).Consistent with the role of the ATR-dependent checkpoint in replication fork stability, common fragile sites, located in late-replicating regions of the genome, are significantly more unstable (5-10-fold) in the absence of ATR or Chk1 (19, 20). Because these sites are favored regions of instability in oncogene-transformed cells and preneoplastic lesions (30, 31), it is possible that the increased tumor incidence observed in ATR haploinsufficient mice (5, 32) may be related to subtle increases in genomic instability. Together, these studies indicate that maintenance of replication fork stability may contribute to tumor suppression.It is important to note that prevention of fork collapse represents an early response to problems occurring during DNA replication. In the event of fork collapse into DSBs, homologous recombination (HR) has also been demonstrated to play a key role in genome stability during S phase by catalyzing recombination between sister chromatids as a means to re-establish replication forks (33). Importantly, a facilitator of homologous recombination, H2AX, has been shown to be phosphorylated under conditions that cause replication fork collapse (18, 34).Phosphorylation of H2AX occurs predominantly upon DSB formation (34-38) and has been reported to require ATM, DNA-PKcs, or ATR, depending on the context (37-42). Although H2AX is not essential for HR, studies have demonstrated that H2AX mutation leads to deficiencies in HR (43, 44), and suppresses events associated with homologous recombination, such as the focal accumulation of Rad51, BRCA1, BRCA2, ubiquitinated-FANCD2, and Ubc13-mediated chromatin ubiquitination (43, 45-51). Therefore, through its contribution to HR, it is possible that H2AX plays an important role in replication fork stability as part of a salvage pathway to reinitiate replication following collapse.If ATR prevents the collapse of stalled replication forks into DSBs, and H2AX facilitates HR-mediated restart, the combined deficiency in ATR and H2AX would be expected to dramatically enhance the accumulation of DSBs upon replication fork stalling. Herein, we utilize both partial and complete elimination of ATR and H2AX to demonstrate that these genes work cooperatively in non-redundant pathways to suppress DSBs during S phase. As discussed, these studies imply that the various components of replication fork protection and regeneration cooperate to maintain replication fork stability. Given the large number of genes involved in each of these processes, it is possible that combined deficiencies in these pathways may be relatively frequent in humans and may synergistically influence the onset of age-related diseases and cancer.  相似文献   

14.
15.
16.
17.
The RIG-I-like receptors (RLRs), RIG-I and MDA5, recognize single-stranded RNA with 5′ triphosphates and double-stranded RNA (dsRNA) to initiate innate antiviral immune responses. LGP2, a homolog of RIG-I and MDA5 that lacks signaling capability, regulates the signaling of the RLRs. To establish the structural basis of dsRNA recognition by the RLRs, we have determined the 2.0-Å resolution crystal structure of human LGP2 C-terminal domain bound to an 8-bp dsRNA. Two LGP2 C-terminal domain molecules bind to the termini of dsRNA with minimal contacts between the protein molecules. Gel filtration chromatography and analytical ultracentrifugation demonstrated that LGP2 binds blunt-ended dsRNA of different lengths, forming complexes with 2:1 stoichiometry. dsRNA with protruding termini bind LGP2 and RIG-I weakly and do not stimulate the activation of RIG-I efficiently in cells. Surprisingly, full-length LGP2 containing mutations that abolish dsRNA binding retained the ability to inhibit RIG-I signaling.The innate immune response is the first line of defense against invading pathogens; it is the ubiquitous system of defense against microbial infections (1). Toll-like receptors (TLRs)3 and RIG-I (retinoic acid-inducible gene 1)-like receptors (RLRs) play key roles in innate immune response toward viral infection (2-5). Toll-like receptors TLR3, TLR7, and TLR8 sense viral RNA released in the endosome following phagocytosis of the pathogens (6). RIG-I-like receptors RIG-I and MDA5 detect viral RNA from replicating viruses in infected cells (3, 7, 8). Stimulation of these receptors leads to the induction of type I interferons (IFNs) and other proinflammatory cytokines, conferring antiviral activity to the host cells and activating the acquired immune responses (4, 9).RIG-I discriminates between viral and host RNA through specific recognition of the uncapped 5′-triphosphate of single-stranded RNA (5′ ppp ssRNA) generated by viral RNA polymerases (10, 11). In addition, RIG-I also recognizes double-stranded RNA generated during RNA virus replication (7, 12). Transfection of cells with synthetic double-stranded RNA stimulates the activation of RIG-I (13, 14). Synthetic dsRNA mimics, such as polyinosinic-polycytidylic acid (poly(I·C)), can activate MDA5 when introduced into the cytoplasm of cells. Digestion of poly(I·C) with RNase III transforms poly(I·C) from a ligand for MDA5 into a ligand for RIG-I, suggesting that MDA5 recognizes long dsRNA, whereas RIG-I recognizes short dsRNA (15). Studies of RIG-I and MDA5 knock-out mice confirmed the essential roles of these receptors in antiviral immune responses and demonstrated that they sense different sets of RNA viruses (12, 16).RIG-I and MDA5 contain two caspase recruiting domains (CARDs) at their N termini, a DEX(D/H) box RNA helicase domain, and a C-terminal regulatory or repressor domain (CTD). The helicase domain and the CTD are responsible for viral RNA binding, whereas the CARDs are required for signaling (3, 8). The current model of RIG-I activation suggests that under resting conditions RIG-I is in a suppressed conformation, and viral RNA binding triggers a conformation change that leads to the exposure of the CARDs for the recruitment of the downstream protein IPS-1 (also known as MAVS, Cardif, or VISA) (14, 17). Limited proteolysis of the RIG-I·dsRNA complex showed that RIG-I residues 792-925 of the CTD are involved in dsRNA and 5′ ppp ssRNA binding (14). The CTD of RIG-I overlaps with the C terminus of the previously identified repressor domain (18). The structures of RIG-I and LGP2 (laboratory of genetics and physiology 2) CTD in isolation have been determined by x-ray crystallography and NMR spectroscopy (14, 19, 20). A large, positively charged surface on RIG-I recognizes the 5′ triphosphate group of viral ssRNA (14, 19). RNA binding studies by titrating RIG-I CTD with dsRNA and 5′ ppp ssRNA suggested that overlapping sets of residues on this charged surface are involved in RNA binding (14). Mutagenesis of several positively charged residues on this surface either reduces or disrupts RNA binding by RIG-I, and these mutations also affect the induction of IFN-β in vivo (14, 19). However, the exact nature of how the RLRs recognize viral RNA and how RNA binding activates these receptors remains to be established.LGP2 is a homolog of RIG-I and MDA5 that lacks the CARDs and thus has no signaling capability (21, 22). The expression of LGP2 is inducible by dsRNA or IFN treatment as well as virus infection (21). Overexpression of LGP2 inhibits Sendai virus and Newcastle disease virus signaling (21). When coexpressed with RIG-I, LGP2 can inhibit RIG-I signaling through the interaction of its CTD with the CARD and the helicase domain of RIG-I (18). LGP2 could suppress RIG-I signaling by three possible ways (23): 1) binding RNA with high affinity, thereby sequestering RNA ligands from RIG-I; 2) interacting directly with RIG-I to block the assembly of the signaling complex; and 3) competing with IKKi (IκB kinase ε) in the NF-κB signaling pathway for a common binding site on IPS-1. To elucidate the structural basis of dsRNA recognition by the RLRs, we have crystallized human LGP2 CTD (residues 541-678) bound to an 8-bp double-stranded RNA and determined the structure of the complex at 2.0 Å resolution. The structure revealed that LGP2 CTD binds to the termini of dsRNA. Mutagenesis and functional studies showed that dsRNA binding is likely not required for the inhibition of RIG-I signaling by LGP2.  相似文献   

18.
RNAs of many positive strand RNA viruses lack a 5′ cap structure and instead rely on cap-independent translation elements (CITEs) to facilitate efficient translation initiation. The mechanisms by which these RNAs recruit ribosomes are poorly understood, and for many viruses the CITE is unknown. Here we identify the first CITE of an umbravirus in the 3′-untranslated region of pea enation mosaic virus RNA 2. Chemical and enzymatic probing of the ∼100-nucleotide PEMV RNA 2 CITE (PTE), and mutagenesis revealed that it forms a long, bulged helix that branches into two short stem-loops, with a possible pseudoknot interaction between a C-rich bulge at the branch point and a G-rich bulge in the main helix. The PTE inhibited translation in trans, and addition of eIF4F, but not eIFiso4F, restored translation. Filter binding assays revealed that the PTE binds eIF4F and its eIF4E subunit with high affinity. Tight binding required an intact cap-binding pocket in eIF4E. Among many PTE mutants, there was a strong correlation between PTE-eIF4E binding affinity and ability to stimulate cap-independent translation. We conclude that the PTE recruits eIF4F by binding eIF4E. The PTE represents a different class of translation enhancer element, as defined by its structure and ability to bind eIF4E in the absence of an m7G cap.Regulation of translation occurs primarily at the initiation step. This involves recognition of the 5′ m7G(5′)ppp(5′)N cap structure on the mRNA by initiation factors, which recruit the ribosome to the 5′-end of the mRNA (15). The 5′ cap structure and the poly(A) tail are necessary for efficient recruitment of initiation factors on eukaryotic mRNAs (3, 68). The cap is recognized by the eIF4E subunit of eukaryotic translation initiation factor complex eIF4F (or the eIFiso4E subunit of eIFiso4F in higher plants). The poly(A) tail is recognized by poly(A)-binding protein. In plants, eIF4F is a heterodimer consisting of eIF4E and eIF4G, the core scaffolding protein to which the other factors bind. eIF4A, an ATPase/RNA helicase, interacts with eIF4F but is not part of the eIF4F heterodimer (9, 10). For translation initiation, the purpose of eIF4E is to bring eIF4G to the capped mRNA. eIF4G then recruits the 43 S ternary ribosomal complex via interaction with eIF3.The RNAs of many positive sense RNA viruses contain a cap-independent translation element (CITE)3 that allows efficient translation in the absence of a 5′ cap structure (1113). In animal viruses and some plant viruses, the CITE is an internal ribosome entry site (IRES) located upstream of the initiation codon. Most viral IRESes neither interact with nor require eIF4E, because they lack the m7GpppN structure, which, until this report, was thought to be necessary for mRNA to bind eIF4E with high affinity (3, 14). Translation initiation efficiency of mRNA is also influenced by the length of, and the degree of secondary structure in the 5′ leader (1517).Many uncapped plant viral RNAs harbor a CITE in the 3′-UTR that confers highly efficient translation initiation at the 5′-end of the mRNA (1822). These 3′ CITEs facilitate ribosome entry and apparently conventional scanning at the 5′-end of the mRNA (17, 23, 24). A variety of unrelated structures has been found to function as 3′ CITEs, suggesting that they recruit the ribosome by different interactions with initiation factors (13).The factors with which a plant CITE interacts to recruit the ribosome have been identified for only a potyvirus, a luteovirus, and a satellite RNA. The 143-nt 5′-UTR CITE of the potyvirus, tobacco etch virus is an IRES that functions by binding of its AU-rich pseudoknot structure with eIF4G (25). It binds eIF4G with up to 30-fold greater affinity than eIFiso4G and does not require eIF4E for IRES activity. In addition to RNA elements, the genome-linked viral protein (VPg) of potyviruses may participate in cap-independent translation initiation by interacting with the eIF4E and eIFiso4E subunits of eIF4F and eIFiso4F, respectively (2631). In contrast, the 130-nt cap-independent translation enhancer domain (TED) in the 3′-UTR of satellite tobacco necrosis virus (STNV) RNA forms a long bulged stem-loop, which interacts strongly with both eIF4F and eIFiso4F and weakly with their eIF4E and eIFiso4E subunits (32), suggesting that the TED requires the full eIF4F or eIFiso4F for a biologically relevant interaction. Barley yellow dwarf luteovirus (BYDV) and several other viruses, have a different structure, called a BYDV-like CITE (BTE), in the 3′-UTR. The BTE is characterized by a 17-nt conserved sequence incorporated in a structure with a variable number of stem-loops radiating from a central junction (20, 33, 34). It requires and binds the eIF4G subunit of eIF4F and does not bind free eIF4E, eIFiso4E, or eIFiso4G, although eIF4E slightly enhances the BTE-eIF4G interaction (35). Other 3′ CITEs have been identified, but the host factors with which they interact are unknown.Here we describe unprecedented factor interactions of a CITE found in an umbravirus and a panicovirus. Umbraviruses show strong similarity to the Luteovirus and Dianthovirus genera in (i) the sequence of the replication genes encoded by ORFs 1 and 2, (ii) the predicted structure of the frameshift signals required for translation of the RNA-dependent RNA polymerase from ORF 2 (36, 37), (iii) the absence of a poly(A) tail, and (iv) the lack of a 5′ cap structure (37, 38). Umbraviruses are unique in that they encode no coat protein. For the umbravirus pea enation mosaic virus 2 (PEMV-2), the coat protein is provided by PEMV-1, an enamovirus (39). Uncapped PEMV-2 RNA (PEMV RNA 2), transcribed in vitro, is infectious in pea (Pisum sativa),4 indicating it must be translated cap-independently. The 3′-UTRs of some umbraviruses such as Tobacco bushy top virus and Groundnut rosette virus harbor sequences resembling BYDV-like CITEs (BTE).5 However, no BTE is apparent in the 3′-UTR of PEMV RNA 2. In this report we identify a different class of CITE in the 705-nt long 3′-UTR of PEMV RNA 2, determine its secondary structure, which may include an unusual pseudoknot, and we show that, unlike any other natural uncapped RNA, it has a high affinity for eIF4E, which is necessary to facilitate cap-independent translation.  相似文献   

19.
RecF pathway proteins play an important role in the restart of stalled replication and DNA repair in prokaryotes. Following DNA damage, RecF, RecR, and RecO initiate homologous recombination (HR) by loading of the RecA recombinase on single-stranded (ss) DNA, protected by ssDNA-binding protein. The specific role of RecF in this process is not well understood. Previous studies have proposed that RecF directs the RecOR complex to boundaries of damaged DNA regions by recognizing single-stranded/double-stranded (ss/ds) DNA junctions. RecF belongs to ABC-type ATPases, which function through an ATP-dependent dimerization. Here, we demonstrate that the RecF of Deinococcus radiodurans interacts with DNA as an ATP-dependent dimer, and that the DNA binding and ATPase activity of RecF depend on both the structure of DNA substrate, and the presence of RecR. We found that RecR interacts as a tetramer with the RecF dimer. RecR increases the RecF affinity to dsDNA without stimulating ATP hydrolysis but destabilizes RecF binding to ssDNA and dimerization, likely due to increasing the ATPase rate. The DNA-dependent binding of RecR to the RecF-DNA complex occurs through specific protein-protein interactions without significant contributions from RecR-DNA interactions. Finally, RecF neither alone nor in complex with RecR preferentially binds to the ss/dsDNA junction. Our data suggest that the specificity of the RecFOR complex toward the boundaries of DNA damaged regions may result from a network of protein-protein and DNA-protein interactions, rather than a simple recognition of the ss/dsDNA junction by RecF.Homologous recombination (HR)2 is one of the primary mechanisms by which cells repair dsDNA breaks (DSBs) and ssDNA gaps (SSGs), and is important for restart of stalled DNA replication (1). HR is initiated when RecA-like recombinases bind to ssDNA forming an extended nucleoprotein filament, referred to as a presynaptic complex (2). The potential for genetic rearrangements dictates that HR initiation is tightly regulated at multiple levels (1). During replication, the ssDNA-binding protein (SSB) protects transiently unwound DNA chains, preventing interactions with recombinases. Following DNA damage, recombination mediator proteins (RMPs) initiate HR by facilitating the formation of the recombinase filaments with ssDNA, while removing SSB (3, 4). Mutations in human proteins involved in HR initiation are linked to cancer predisposition, chromosome instability, UV sensitivity, and premature aging diseases (48). To date, little is known about the mechanism by which RMPs regulate the formation of the recombinase filaments on the SSB-protected ssDNA.In Escherichia coli, there are two major recombination pathways, RecBCD and RecF (9, 10). A helicase/nuclease RecBCD complex processes DSBs and recruits RecA on ssDNA in a sequence-specific manner (1113). The principle players in the RecF pathway are the RecF, RecO, and RecR proteins, which form an epistatic group that is important for SSG repair, for restart of stalled DNA replication, and under specific conditions, can also process DSBs (1420). Homologs of RecF, -O, and -R are present in the majority of known bacteria (21), including Deinococcus radiodurans, extremely radiation-resistant bacteria that lacks the RecBCD pathway, yet is capable of repairing thousands of DSBs (22, 23). In addition, the sequence or functional homologs of RecF pathway proteins are involved in similar pathways in eukaryotes that include among others WRN, BLM, RAD52, and BRCA2 proteins (48).The involvement of all three RecF, -O, and -R proteins in HR initiation is well documented by genetic and cellular approaches (18, 2430), yet their biochemical functions in the initiation process remain unclear, particularly with respect to RecF. RecO and RecR proteins are sufficient to promote formation of the RecA filament on SSB-bound ssDNA in vitro (27). The UV-sensitive phenotype of recF mutants can be suppressed by RecOR overexpression, suggesting that RecF may direct the RMP complex to DNA-damaged regions where HR initiation is required (31). In agreement with this hypothesis, RecF dramatically increases the efficiency of the RecA loading at ds/ssDNA junctions with a 3′ ssDNA extension under specific conditions (32). RecF and RecR proteins also prevent the RecA filaments from extending into dsDNA regions adjacent to SSGs (33). These data suggest that RecF may directly recognize an ss/dsDNA junction structure (34). However, DNA binding experiments have not provided clear evidence to support such a hypothesis (11).The targeting promoted by RecF may also occur through more complex processes. RecF shares a high structural similarity with the head domain of Rad50, an ABC-type ATPase that recognizes DSBs and initiates repair in archaea and eukaryotes (35). All known ABC-type ATPases function as oligomeric complexes in which a sequence of inter- and intra-molecular interactions is triggered by the ATP-dependent dimerization and the dimer-dependent ATP hydrolysis (3639). RecF is also an ATP-dependent DNA-binding protein and a weak DNA-dependent ATPase (11, 40). RecF forms an ATP-dependent dimer and all three conserved motifs (Walker A, Walker B, and “signature”) of RecF are important for ATP-dependent dimerization, ATP hydrolysis, and functional resistance to DNA damage (35). Thus, RecF may function in recombination initiation through a complex pathway of protein-protein and DNA-protein interactions regulated by ATP-dependent RecF dimerization.In this report, we present a detailed characterization of the RecF dimerization, and its role in the RecF interaction with various DNA substrates, with RecR, and in ATP hydrolysis. Our data outline the following key findings. First, RecF interacts with DNA as a dimer. Second, neither RecF alone nor the RecFR complex preferentially binds the ss/dsDNA junction. Finally, RecR changes the ATPase activity and the DNA binding of RecF by destabilizing the interaction with ssDNA, and greatly enhancing the interaction with dsDNA. Our results suggest that the specificity of RecF for the boundaries of SSGs is likely to result from a sequence of protein-protein interaction events rather than a simple RecF ss/dsDNA binding, underlining a highly regulated mechanism of the HR initiation by the RecFOR proteins.  相似文献   

20.
Persistent infection with hepatitis C virus (HCV) is a major cause of chronic liver diseases. The aim of this study was to identify host cell factor(s) participating in the HCV replication complex (RC) and to clarify the regulatory mechanisms of viral genome replication dependent on the host-derived factor(s) identified. By comparative proteome analysis of RC-rich membrane fractions and subsequent gene silencing mediated by RNA interference, we identified several candidates for RC components involved in HCV replication. We found that one of these candidates, creatine kinase B (CKB), a key ATP-generating enzyme that regulates ATP in subcellular compartments of nonmuscle cells, is important for efficient replication of the HCV genome and propagation of infectious virus. CKB interacts with HCV NS4A protein and forms a complex with NS3-4A, which possesses multiple enzyme activities. CKB upregulates both NS3-4A-mediated unwinding of RNA and DNA in vitro and replicase activity in permeabilized HCV replicating cells. Our results support a model in which recruitment of CKB to the HCV RC compartment, which has high and fluctuating energy demands, through its interaction with NS4A is important for efficient replication of the viral genome. The CKB-NS4A association is a potential target for the development of a new type of antiviral therapeutic strategy.Hepatitis C virus (HCV) infection represents a significant global healthcare burden, and current estimates suggest that a minimum of 3% of the world''s population is chronically infected (4, 19). The virus is responsible for many cases of severe chronic liver diseases, including cirrhosis and hepatocellular carcinoma (4, 16, 19). HCV is a positive-stranded RNA virus belonging to the family Flaviviridae. Its ∼9.6-kb genome is translated into a single polypeptide of about 3,000 amino acids (aa), in which the nonstructural (NS) proteins NS2, NS3, NS4A, NS4B, NS5A, and NS5B reside in the C-terminal half region (6, 34, 44). NS4A, a small 7-kDa protein, functions as a cofactor for NS3 to enhance NS3 enzyme activities such as serine protease and helicase activities. The hydrophobic N-terminal region of NS4A, which is predicted to form a transmembrane α-helix, is responsible for membrane anchorage of the NS3-4A complex (8, 44, 50), and the central region of NS4A is important for the interaction with NS3 (10, 44). A recent study demonstrated the involvement of the C terminus of NS4A in the regulation of NS5A hyperphosphorylation and viral replication (28).The development of HCV replicon technology several years ago accelerated research on viral RNA replication (7, 44). Furthermore, a robust cell culture system for propagation of infectious HCV particles was developed using a viral genome of HCV genotype 2a, JFH-1 strain, enabling us to study every process in the viral life cycle (27, 47, 54). RNA derived from genotype 1a, HCV H77, containing cell-culture adaptive mutations, also produces infectious viruses (52). Using these systems, it has been reported that the HCV genome replicates in a distinct, subcellular replication complex (RC) compartment, which includes NS3-5B and the viral RNA (2, 14, 33). The RC forms in a distinct compartment with high concentrations of viral and cellular components located on detergent-resistant membrane (DRM) structures, possibly a lipid-raft structure (2, 41), which may protect the RC from external proteases and nucleases. Almost all processes in viral replication are dependent on the host cell''s machinery and involve intimate interaction between viral and host proteins. However, the functional roles of host factors interacting with the HCV RC in viral genome replication remain ambiguous.To gain a better understanding of cellular factors that are components of the HCV RC and that function as regulators of viral replication, a comparative proteomic analysis of DRM fractions from HCV replicon and parental cells and subsequent RNA interference (RNAi) silencing of selected genes were performed. We identified creatine kinase B (CKB) as a key factor for the HCV genome replication. CKB catalyzes the reversible transfer of the phosphate group of phosphocreatine (pCr) to ADP to yield ATP and creatine and is known to play important roles in local delivery and cellular compartmentalization of ATP (48, 51). The findings obtained here suggest that recruitment of CKB to the HCV RC, through CKB interaction with NS4A, is essential for maintenance or enhancement of viral replicase activity.  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