首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Dynamic interactions between components of the outer (OM) and inner (IM) membranes control a number of critical mitochondrial functions such as channeling of metabolites and coordinated fission and fusion. We identify here the mitochondrial AAA+ ATPase protein ATAD3A specific to multicellular eukaryotes as a participant in these interactions. The N-terminal domain interacts with the OM. A central transmembrane segment (TMS) anchors the protein in the IM and positions the C-terminal AAA+ ATPase domain in the matrix. Invalidation studies in Drosophila and in a human steroidogenic cell line showed that ATAD3A is required for normal cell growth and cholesterol channeling at contact sites. Using dominant-negative mutants, including a defective ATP-binding mutant and a truncated 50-amino-acid N-terminus mutant, we showed that ATAD3A regulates dynamic interactions between the mitochondrial OM and IM sensed by the cell fission machinery. The capacity of ATAD3A to impact essential mitochondrial functions and organization suggests that it possesses unique properties in regulating mitochondrial dynamics and cellular functions in multicellular organisms.Mitochondria not only supply cells with the bulk of their ATP but also contribute to the fine regulation of metabolism, calcium homeostasis, and apoptosis (27). Coordination of these functions is dependent on the dynamic nature of mitochondria (5). These organelles constantly fuse and divide to form small spheres, short rods, or long tubules and are actively transported to specific subcellular locations. These processes are essential for mammalian development, and defects can lead to degenerative diseases and cancers (9, 17). In eukaryotes, these organellar gymnastics are controlled by numerous pathways that preserve proper mitochondrial morphology and function (30, 45). The best-understood mitochondrial process is the fusion and fission pathways, which rely on conserved GTPases, and their binding partners to regulate organelle connectivity (10, 18, 45). There are also evidences that dynamic interactions between the outer membrane (OM) and inner membrane (IM) exist for coordinated fusion and fission, channeling of metabolites, and protein transport, but proteins playing a role in these interactions have yet to be identified (34). In the present study, we provide a detailed biochemical and functional characterization of the mitochondrial AAA+ ATPase ATAD3A protein that is present exclusively in multicellular eukaryotes and which participates in the control of mitochondrial dynamics at the interface between the IMs and OMs. Proteins related to the Atad3A genes have been previously identified in proteomic surveys of mouse brain mitochondria (28) and liver mitochondrial inner membrane (8), as mitochondrial DNA-binding proteins (4, 21, 44) and as nuclear mRNA-associated proteins (6). The Atad3A protein has also been identified as a cell surface antigen in some human tumors (16). Functional genomics identified the Drosophila Atad3A ortholog (bor) as a major gene positively regulated by the TOR (for target of rapamycin) signaling pathway involved in cell growth and division (19). In our laboratory, we identified ATAD3A as a specific target for the Ca2+/Zn2+-binding S100B protein (B. Gilquin et al., unpublished data). We here show that ATAD3A is anchored into the mitochondrial IM at contact sites with the OM. The N-terminal domain of ATAD3A interacts with the inner surface of the OM and its C-terminal AAA ATPase domain localizes in a specific matrix compartment. Thanks to its simultaneous interaction with two membranes, ATAD3A regulates mitochondrial dynamics at the interface between the IMs and OMs and controls diverse cell responses ranging from cell growth, channeling of cholesterol, and mitochondrial fission.  相似文献   

2.
3.
Glutathione constitutes a key player in the thiol redox buffer in many organisms. However, the gram-positive bacteria Bacillus subtilis and Staphylococcus aureus lack this low-molecular-weight thiol. Recently, we identified S-cysteinylated proteins in B. subtilis after treatment of cells with the disulfide-generating electrophile diamide. S cysteinylation is thought to protect protein thiols against irreversible oxidation to sulfinic and sulfonic acids. Here we show that S thiolation occurs also in S. aureus proteins after exposure to diamide. We further analyzed the formation of inter- and intramolecular disulfide bonds in cytoplasmic proteins using diagonal nonreducing/reducing sodium dodecyl sulfate gel electrophoresis. However, only a few proteins were identified that form inter- or intramolecular disulfide bonds under control and diamide stress conditions in B. subtilis and S. aureus. Depletion of the cysteine pool was concomitantly measured in B. subtilis using a metabolomics approach. Thus, the majority of reversible thiol modifications that were previously detected by two-dimensional gel fluorescence-based thiol modification assay are most likely based on S thiolations. Finally, we found that a glutathione-producing B. subtilis strain which expresses the Listeria monocytogenes gshF gene did not show enhanced oxidative stress resistance compared to the wild type.Cysteine thiols in proteins fulfill an important and diverse set of cellular functions. In particular, they participate in enzymatic catalysis; in metal coordination, such as in the generation of Fe-S-clusters; and in determining the spatial structure of proteins via disulfide bond formation (3, 22, 23, 38). Cysteines are strong nucleophiles amenable to posttranslational modifications by reactive oxygen species (ROS) and reactive nitrogen species, leading to disulfides; to sulfenic, sulfinic, or sulfonic acids; mixed disulfides with low-molecular-weight (LMW) thiols (S thiolations); and S nitrosylations (7, 16, 17, 27).The redox status of the cytoplasm is under physiological conditions in a reduced state. Thus, most cysteines are present as free thiols (6). Because aerobic organisms have to cope with oxidative stress caused by ROS, such as superoxide anions, hydrogen peroxide, or hydroxyl radicals, they need to employ effective mechanisms that maintain the reduced state. In gram-negative bacteria, the thiol-disulfide balance is accomplished by the glutathione (GSH) system, a thiol-based redox buffer. The GSH system consists of glutaredoxin (Grx), GSH (γ-glutamylcysteinyl glycine), GSH reductase, and GSH peroxidase (34). Reduction of disulfides occurs via sequential electron transfer from glutaredoxin and reduced GSH; oxidized GSH (GSSG) is reduced by the NADPH-dependent GSH reductase. GSH peroxidase enables the direct detoxification of ROS by GSH oxidation.However, many gram-positive bacteria lack genes for GSH biosynthesis. Actinomycetes instead use a thiol redox buffer based on mycothiol (50). Bacillus subtilis, Staphylococcus aureus, and other gram-positive bacteria rely on different thiol redox buffers based on cysteine, the novel 398-Da bacillithiol (BSH), or coenzyme A (CoA) (15, 52). To maintain the reduced state of the cytoplasm, most bacteria use enzymatic systems for disulfide bond reduction such as the thioredoxin (Trx) system, which is highly conserved in gram-negative bacteria (3, 10). The Trx system consists of thioredoxin (TrxA) and the NADPH-dependent thioredoxin reductase (TrxB).Any imbalance in the cellular redox state caused by ROS elicits expression of a repertoire of different proteins, commonly under the control of a redox-sensing regulator: for example, OxyR in Escherichia coli and PerR, OhrR, SarZ, and Spx in B. subtilis and S. aureus, respectively (11, 12, 41, 55, 58, 64-66). The subsequently induced proteins detoxify ROS and restore and protect the normal physiological redox state in the cell.Besides ROS and reactive nitrogen species, so-called “reactive electrophilic species” (RES) affect the thiol redox balance. RES include different chemical compounds such as aldehydes, quinones, and the azo compound diamide (2, 43, 45, 46, 53, 66). Quinones and aldehydes have electron-deficient centers that result in thiol-(S) alkylation of cysteine. Exposure of cells to diamide induces the oxidative as well as the electrophile stress response in B. subtilis (43, 45, 53). The toxicity of diamide is based on disulfide bond formation (40), which was recently visualized in B. subtilis and S. aureus by the fluorescence alkylation of oxidized thiols (FALKO) assay (32, 64). It was thought that the formation of nonnative inter- and intramolecular disulfide bonds results in damage of proteins.However, more recent findings demonstrate that diamide stress leads also to S thiolations: formation of disulfide bonds between proteins and LMW thiols (8, 13, 33). S thiolations prevent protein thiols from irreversible oxidation to sulfinic and sulfonic acids, but also affect enzyme activity (35, 47) and signal transduction (39, 42). In B. subtilis, we have identified a few cytoplasmic proteins that are S cysteinylated (33). In addition, the organic peroxide sensor OhrR was inactivated by an S bacillithiolation in B. subtilis (42).Cysteine, BSH, and CoA are also dominant LMW thiols in S. aureus (52). In this study, we have investigated in more detail the extents of S thiolations and inter- and intramolecular disulfide bond formation of B. subtilis and S. aureus in response to disulfide stress. The results showed that exposure to diamide leads to S thiolations in S. aureus. Using a nonreducing/reducing sodium dodecyl sulfate (SDS) diagonal electrophoresis approach, proteins with intermolecular disulfide bonds could be distinguished from proteins with intramolecular disulfide bonds (57). The results support that the majority of reversible thiol oxidations are based on S thiolations rather than disulfide bonds between proteins. Depletion of the free cysteine pool in B. subtilis after exposure to diamide supports this finding. To assess if GSH may have a bearing on the thiol redox buffer of B. subtilis, the gshF gene of Listeria monocytogenes (gshFLm) was expressed in B. subtilis, enabling GSH biosynthesis (29). Although GSH production does not enhance the resistance to oxidative stress in B. subtilis, it participates in the formation of S thiolations.  相似文献   

4.
5.
6.
The Ysa type III secretion (T3S) system enhances gastrointestinal infection by Yersinia enterocolitica bv. 1B. One effector protein targeted into host cells is YspP, a protein tyrosine phosphatase. It was determined in this study that the secretion of YspP requires a chaperone, SycP. Genetic analysis showed that deletion of sycP completely abolished the secretion of YspP without affecting the secretion of other Ysps by the Ysa T3S system. Analysis of the secretion and translocation signals of YspP defined the first 73 amino acids to form the minimal region of YspP necessary to promote secretion and translocation by the Ysa T3S system. Function of the YspP secretion/translocation signals was dependent on SycP. Curiously, when YspP was constitutively expressed in Y. enterocolitica bv. 1B, it was recognized and secreted by the Ysc T3S system and the flagellar T3S system. In these cases, the first 21 amino acids were sufficient to promote secretion, and while SycP did enhance secretion, it was not essential. However, neither the Ysc T3S system nor the flagellar T3S system translocated YspP into mammalian cells. This supports a model where SycP confers secretion/translocation specificities for YspP by the Ysa T3S system. A series of biochemical approaches further established that SycP specifically interacts with YspP and protected YspP degradation in the cell prior to secretion. Collectively, the evidence suggests that YspP secretion by the Ysa T3S system is a posttranslational event.Many gram-negative bacteria have evolved sophisticated delivery systems termed type III secretion (T3S) systems to transport effector proteins into the cytosols of eukaryotic host cells (10, 21, 22). The translocated effectors manipulate host cell activities in various ways, thereby permitting the establishment of a pathogenic or symbiotic interaction (20). T3S systems are ancestrally related to the flagellar T3S system, having in common a basal body spanning the inner and outer bacterial membranes responsible for the appropriate selection of polypeptides delivered into a hollow channel leading out of the bacterium. At the outer surface, flagellar polypeptides travel the length of the adjoining hook and filament, but in T3S systems, the secreted polypeptides pass through a special hollow needle that extends away from the bacterium to the targeted host cell (10, 21, 22). Heterologous multimeric proteins localized to the tip of the needle form the translocon, a porelike channel that is assembled in the eukaryotic plasma membrane, enabling the injection of bacterial effectors (24, 48, 51).Two terminologies are distinctly used to describe protein transport by T3S systems. While “secretion” is a transport event for proteins from the bacterial cytosol into the extracellular milieu, “translocation” is a transport event for proteins from the bacterial cytosol into the eukaryotic host''s cytosol. Generally, secretion but not translocation is mediated by the first 20 amino acids of effector proteins (41, 46, 47), albeit mRNA sequences at the N terminus of some proteins have been also considered to function as the secretion signals (3, 44). This secretion event is independent of the presence of cognate effector chaperones (46, 59). Despite no conservation of the amino acids among the secretion signals, amphipathic or disordered secondary structures of the peptides are thought to function as the secretion signals recognized by the T3S apparatuses (22, 34, 35). In contrast, translocation usually requires both the secretion (the first 20 amino acids) and the translocation (amino acids 20 to 100) signals (46, 47, 59). This translocation event is efficiently mediated by the presence of the cognate chaperones (9, 14, 30), and the chaperone-effector complexes have been proposed to function as the three-dimensional signals recognized by the T3S apparatuses (5, 33, 38, 49, 50).Many T3S effectors employ cognate chaperones in the bacterial cytoplasm (43, 57). The effector chaperones have been categorized into two subgroups, class 1A and class 1B, primarily based on the substrate properties (and the gene locations) (13, 43). Class 1A chaperones commonly bind to one effector, and most of them are encoded by genes located adjacent to the gene encoding the cognate effectors. In contrast, class 1B chaperones bind to multiple effectors and are encoded by genes located within operons that code for structural components of the T3S apparatus that are distant to the cognate effector genes. Evolutionally, this subgroup of chaperones is thought to be an archetype of effector chaperones. Although T3S effector chaperones lack primary sequence similarity even in same subgroup, overall the effector chaperones whose three-dimensional structures are solved share similar folds, consisting of three α-helices and five β-strands (5, 36, 38, 49, 54). Similarly, effector chaperones share the common biochemical characteristics of acidic properties (pI 4 to 5) and low molecular masses (12 to 15 kDa), with a tendency to form homodimers (43). These homodimers recognize the chaperone binding domains (CBD) of the cognate effectors, which are usually located in the amino-terminal 20 to 100 amino acids (translocation signal) of the effector (19, 30, 59). Despite the wealth of information about individual chaperones, a universally accepted model for the mechanisms by which they promote secretion is lacking. One study shows that the guidance of chaperone-effector complexes toward the T3S apparatus is provided by the affinity of their chaperones to the ATPase of the T3S apparatus, whereby the ATPase releases the chaperones from the complexes and then unfolds the cognate effector for secretion (2). Several additional functions of T3S effector chaperones have been reported, including the prevention of effector aggregation prior to delivery to the secretion system, limitation of premature interactions, and protection of effectors from protease degradation in bacterial cells (17, 43). When an organism has multiple T3S pathways, as is the case for some Yersinia spp., there is the opportunity to gain new insight into how a given chaperone might influence T3S system specificity for substrates. Without direct testing of the aforementioned mechanistic models, the role of a chaperone in T3S and how it affects the overall sequence of pathogenic events is, at best, a conjecture.Highly virulent strains of Yersinia enterocolitica bv. 1B have a total of three T3S systems. The first T3S system (Ysc) is encoded by the virulence plasmid, and it secretes six effectors termed Yops. Ysc T3S is important for systemic infection (11, 12, 42). This T3S system is common to all Yersinia species pathogenic to humans, including another enteropathogen, Yersinia pseudotuberculosis, and the plague pathogen Yersinia pestis. The second system (Ysa) is encoded by a cluster of genes mapping to the Ysa pathogenicity island (25, 53). The Ysa T3S system secretes a set of eight effectors termed Ysps and, interestingly, also secretes three Yops, YopE, YopN, and YopP/YopJ (39, 58, 61). This Ysa T3S system is restricted to clinical isolates of Y. enterocolitica bv. 1B and promotes the initial establishment of infection in gastrointestinal tissue (39, 55). The third T3S system is an integral part of the flagellum and secretes proteins termed Fops to the extracellular milieu (64).Previously, we identified the suite of Ysp proteins secreted by the Ysa T3S system (39). However, little is known about the detailed mechanism by which these proteins are secreted and translocated by this system. Among the Ysp proteins identified, YspP is a protein tyrosine phosphatase (PTPase) whose activity is required for full virulence (39). Here, we found a small open reading frame (ORF) immediately downstream of yspP and designated it sycP. The SycP protein was demonstrated to be a YspP-specific chaperone essential for both the secretion and the translocation of YspP by the Ysa T3S system. In addition, we also examined the secretion specificity requirements for YspP secretion by three different T3S systems as model cases. Interestingly, our data suggest that the mechanisms by which the secretion and translocation signals are recognized are different, depending on the type of T3S system examined.  相似文献   

7.
The sequestration of iron by mammalian hosts represents a significant obstacle to the establishment of a bacterial infection. In response, pathogenic bacteria have evolved mechanisms to acquire iron from host heme. Bacillus anthracis, the causative agent of anthrax, utilizes secreted hemophores to scavenge heme from host hemoglobin, thereby facilitating iron acquisition from extracellular heme pools and delivery to iron-regulated surface determinant (Isd) proteins covalently attached to the cell wall. However, several Gram-positive pathogens, including B. anthracis, contain genes that encode near iron transporter (NEAT) proteins that are genomically distant from the genetically linked Isd locus. NEAT domains are protein modules that partake in several functions related to heme transport, including binding heme and hemoglobin. This finding raises interesting questions concerning the relative role of these NEAT proteins, relative to hemophores and the Isd system, in iron uptake. Here, we present evidence that a B. anthracis S-layer homology (SLH) protein harboring a NEAT domain binds and directionally transfers heme to the Isd system via the cell wall protein IsdC. This finding suggests that the Isd system can receive heme from multiple inputs and may reflect an adaptation of B. anthracis to changing iron reservoirs during an infection. Understanding the mechanism of heme uptake in pathogenic bacteria is important for the development of novel therapeutics to prevent and treat bacterial infections.Pathogenic bacteria need to acquire iron to survive in mammalian hosts (12). However, the host sequesters most iron in the porphyrin heme, and heme itself is often bound to proteins such as hemoglobin (14, 28, 85). Circulating hemoglobin can serve as a source of heme-iron for replicating bacteria in infected hosts, but the precise mechanisms of heme extraction, transport, and assimilation remain unclear (25, 46, 79, 86). An understanding of how bacterial pathogens import heme will lead to the development of new anti-infectives that inhibit heme uptake, thereby preventing or treating infections caused by these bacteria (47, 68).The mechanisms of transport of biological molecules into a bacterial cell are influenced by the compositional, structural, and topological makeup of the cell envelope. Gram-negative bacteria utilize specific proteins to transport heme through the outer membrane, periplasm, and inner membrane (83, 84). Instead of an outer membrane and periplasm, Gram-positive bacteria contain a thick cell wall (59, 60). Proteins covalently anchored to the cell wall provide a functional link between extracellular heme reservoirs and intracellular iron utilization pathways (46). In addition, several Gram-positive and Gram-negative bacterial genera also contain an outermost structure termed the S (surface)-layer (75). The S-layer is a crystalline array of protein that surrounds the bacterial cell and may serve a multitude of functions, including maintenance of cell architecture and protection from host immune components (6, 7, 18, 19, 56). In bacterial pathogens that manifest an S-layer, the “force field” function of this structure raises questions concerning how small molecules such as heme can be successfully passed from the extracellular milieu to cell wall proteins for delivery into the cell cytoplasm.Bacillus anthracis is a Gram-positive, spore-forming bacterium that is the etiological agent of anthrax disease (30, 33). The life cycle of B. anthracis begins after a phagocytosed spore germinates into a vegetative cell inside a mammalian host (2, 40, 69, 78). Virulence determinants produced by the vegetative cells facilitate bacterial growth, dissemination to major organ systems, and eventually host death (76-78). The release of aerosolized spores into areas with large concentrations of people is a serious public health concern (30).Heme acquisition in B. anthracis is mediated by the action of IsdX1 and IsdX2, two extracellular hemophores that extract heme from host hemoglobin and deliver the iron-porphyrin to cell wall-localized IsdC (21, 45). Both IsdX1 and IsdX2 harbor near iron transporter domains (NEATs), a conserved protein module found in Gram-positive bacteria that mediates heme uptake from hemoglobin and contributes to bacterial pathogenesis upon infection (3, 8, 21, 31, 44, 46, 49, 50, 67, 81, 86). Hypothesizing that B. anthracis may contain additional mechanisms for heme transport, we provide evidence that B. anthracis S-layer protein K (BslK), an S-layer homology (SLH) and NEAT protein (32, 43), is surface localized and binds and transfers heme to IsdC in a rapid, contact-dependent manner. These results suggest that the Isd system is not a self-contained conduit for heme trafficking and imply that there is functional cross talk between differentially localized NEAT proteins to promote heme uptake during infection.  相似文献   

8.
Filopodia are dynamic structures found at the leading edges of most migrating cells. IRSp53 plays a role in filopodium dynamics by coupling actin elongation with membrane protrusion. IRSp53 is a Cdc42 effector protein that contains an N-terminal inverse-BAR (Bin-amphipysin-Rvs) domain (IRSp53/MIM homology domain [IMD]) and an internal SH3 domain that associates with actin regulatory proteins, including Eps8. We demonstrate that the SH3 domain functions to localize IRSp53 to lamellipodia and that IRSp53 mutated in its SH3 domain fails to induce filopodia. Through SH3 domain-swapping experiments, we show that the related IRTKS SH3 domain is not functional in lamellipodial localization. IRSp53 binds to 14-3-3 after phosphorylation in a region that lies between the CRIB and SH3 domains. This association inhibits binding of the IRSp53 SH3 domain to proteins such as WAVE2 and Eps8 and also prevents Cdc42-GTP interaction. The antagonism is achieved by phosphorylation of two related 14-3-3 binding sites at T340 and T360. In the absence of phosphorylation at these sites, filopodium lifetimes in cells expressing exogenous IRSp53 are extended. Our work does not conform to current views that the inverse-BAR domain or Cdc42 controls IRSp53 localization but provides an alternative model of how IRSp53 is recruited (and released) to carry out its functions at lamellipodia and filopodia.The ability of a cell to rapidly respond to extracellular cues and direct cytoskeletal rearrangements is dependent on an array of signaling complexes that control actin assembly (58). The protrusive structures at the leading edges of motile cells are broadly defined as lamellipodia or filopodia (14). Lamellae are sheet-like protrusions composed of dendritic actin arrays that drive membrane expansion, with the “lamellipodium” representing a narrow region at the edge of the cell (in culture) characterized by rapid actin polymerization. This F-actin assembly is suggested to require Arp2/3 activity that nucleates new actin filaments from the sides of existing ones (58, 71) and capping proteins that limit the length of these new filaments and stabilize them (7). Arp2/3 activity in turn is regulated by the WASP/WAVE family of proteins, such as N-WASP and WAVE2 (68), whose regulation is a subject of intense interest (12, 29, 36, 41, 56, 76).Filopodia contain parallel bundles of actin filaments containing fascin (22). These are dynamic structures that emanate from the periphery of the cell and are retracted, with occasional attachment (to the dish in culture). Thus, they have been thought to have a sensory or exploratory role during cell migration (28). This is the case for neuronal growth cones, where filopodia sense attractant or repulsive cues and dictate direction in axonal path finding (9, 17, 25, 35). Filopodia have been shown to be important in the context of dendritic-spine development (64, 77), epithelial-sheet closure (26, 60, 79), and cell invasion/metastasis (80, 83).Lamellipodia have been well characterized since the pioneering work of Abercrombie et al. in the early 1970s (2, 3, 4). Filopodia require symmetry breaking at the leading edge (initiation), followed by elongation driven by a filopodial-tip protein complex (14, 28). A few proteins have been identified in this complex; Mena/Vasp serve to prevent capping at the barbed ends of bundled actin filaments (7, 53), and Dia2 promotes F-actin elongation (57, 85). Termination of filopodial elongation is not understood but nonetheless is likely to be tightly regulated. In the absence of F-actin elongation, retraction of the filopodium takes place by a rearward flow of F-actin and filament depolymerization (22).IRSp53 is in a position to play a pivotal role in generating filopodia; this brain-enriched protein was discovered as a substrate of the insulin receptor (87). Subsequently, IRSp53 was identified as an effector for Rac1 (50) and Cdc42 (27, 38), where it participates in filopodium and lamellipodium production (38, 51, 54, 86), neurite extension (27), dendritic-spine morphogenesis (1, 15, 66, 67), cell motility and invasiveness (24). The N terminus of IRSp53 contains a conserved helical domain that is found in five different gene products and is referred to as the IRSp53/MIM homology domain (IMD) (51, 70). This domain has been postulated to bind to Rac1 (50, 70) in a nucleotide-independent manner (52), but no convincing effector-like region has been identified. A Cdc42-specific CRIB-like sequence that does not bind Rac1 (27, 38) allows coupling of this and perhaps related Rho GTPases. The structure of the IMD reveals a zeppelin-shaped dimer that could bind “bent” membranes; thus, its potential as an F-actin-bundling domain (51, 82) could be an in vitro artifact often attributed to proteins with basic patches (46). Although there are reports of F-actin binding at physiological ionic strength (ca. 100 mM KCl) (82, 19), this region when expressed in isolation does not decorate F-actin in vivo.Two reports showed the IMD to be an “inverse-BAR” domain. BAR (Bin-amphipysin-Rvs) domains are found in proteins involved in endocytic trafficking, such as amphipysin and endophilin, and stabilize positively bent membranes, such as those on endocytic vesicles (31, 47). The IMD domains of both IRSp53 (70) and MIM-B (46) associate with lipids and can induce tubulations of PI(3,4,5)P3 or PI(4,5)P2-rich membranes, respectively. These tubulations are equivalent to membrane protrusions and are also referred to as negatively bent membranes. Ectopic expression of the IMD from IRSp53 (51, 70, 82, 86) or two other family members, MIM-B (11, 46) and IRTKS (52), can give rise to cells with many peripheral extensions. MIM-B is said to stimulate lamellipodia (11), while IRTKS generates “short actin clusters” at the cell periphery (52).In IRSp53 is a CRIB-like motif that mediates binding to Cdc42 (27, 38), but the function of this interaction in unclear. Cdc42 could relieve IRSp53 autoinhibition as described for N-Wasp (38), but there is little evidence for this. It has been suggested that Cdc42 controls IRSp53 localization and actin remodeling (27, 38), but another study indicated that these events are Cdc42 independent (19). IRSp53 contains a central SH3 domain that may bind proline-rich proteins, such as Dia1 (23), Mena (38), WAVE2 (49, 50, 69), and Eps8 (19, 24). However, it seems unlikely that all of these represent bona fide partners, and side-by-side comparison is provided in this study. Mena is involved in filopodium production (37), Dia1 in stress fiber formation (81), and WAVE2 in lamellipodium extension (72). Thus, Mena is a better candidate as a partner for IRSp53-mediated filopodia than Dia1 or WAVE2.There is good evidence for IRSp53 as a cellular partner for Eps8 (19). Eps8 is an adaptor protein containing an N-terminal PTB domain that can associate with receptor tyrosine kinases (65), and perhaps β integrins (13), and a C-terminal SH3 domain that can associate with Abi1 (30). Binding of the general adaptor Abi1 appears to positively regulate the actin-capping domain at the C terminus of Eps8 (18). It has been suggested that IRSp53 and Eps8 as a complex regulate cell motility, and perhaps Rac1 activation, via SOS (24); more recently, their roles in filopodium formation have been addressed (19). The involvement of IRSp53, but not MIM-B or IRTKS, in filopodium formation might be related to its role as a Cdc42 effector. We show here that, surprisingly, the CRIB motif is not essential for this activity, but rather, the ability of IRSp53 to associate via its SH3 domain is required, and that this domain is controlled by 14-3-3 binding.We have focused on the regulation of Cdc42 effectors that bind 14-3-3, including IRSp53 and PAK4, which are found as 14-3-3 targets in various proteomic projects (32, 44). In this study, we characterize the binding of 14-3-3 to IRSp53 and uncover how this activity regulates IRSp53 function. The phosphorylation-dependent 14-3-3 binding is GSK3β dependent, and 14-3-3 blocks the accessibility of both the CRIB and SH3 domains of IRSp53, thus indicating its primary function in controlling IRSp53 partners. This regulation of the SH3 domain by 14-3-3 is critical in the proper localization and termination of IRSp53 function to promote filopodium dynamics.  相似文献   

9.
Analysis of Lyme borreliosis (LB) spirochetes, using a novel multilocus sequence analysis scheme, revealed that OspA serotype 4 strains (a rodent-associated ecotype) of Borrelia garinii were sufficiently genetically distinct from bird-associated B. garinii strains to deserve species status. We suggest that OspA serotype 4 strains be raised to species status and named Borrelia bavariensis sp. nov. The rooted phylogenetic trees provide novel insights into the evolutionary history of LB spirochetes.Multilocus sequence typing (MLST) and multilocus sequence analysis (MLSA) have been shown to be powerful and pragmatic molecular methods for typing large numbers of microbial strains for population genetics studies, delineation of species, and assignment of strains to defined bacterial species (4, 13, 27, 40, 44). To date, MLST/MLSA schemes have been applied only to a few vector-borne microbial populations (1, 6, 30, 37, 40, 41, 47).Lyme borreliosis (LB) spirochetes comprise a diverse group of zoonotic bacteria which are transmitted among vertebrate hosts by ixodid (hard) ticks. The most common agents of human LB are Borrelia burgdorferi (sensu stricto), Borrelia afzelii, Borrelia garinii, Borrelia lusitaniae, and Borrelia spielmanii (7, 8, 12, 35). To date, 15 species have been named within the group of LB spirochetes (6, 31, 32, 37, 38, 41). While several of these LB species have been delineated using whole DNA-DNA hybridization (3, 20, 33), most ecological or epidemiological studies have been using single loci (5, 9-11, 29, 34, 36, 38, 42, 51, 53). Although some of these loci have been convenient for species assignment of strains or to address particular epidemiological questions, they may be unsuitable to resolve evolutionary relationships among LB species, because it is not possible to define any outgroup. For example, both the 5S-23S intergenic spacer (5S-23S IGS) and the gene encoding the outer surface protein A (ospA) are present only in LB spirochete genomes (36, 43). The advantage of using appropriate housekeeping genes of LB group spirochetes is that phylogenetic trees can be rooted with sequences of relapsing fever spirochetes. This renders the data amenable to detailed evolutionary studies of LB spirochetes.LB group spirochetes differ remarkably in their patterns and levels of host association, which are likely to affect their population structures (22, 24, 46, 48). Of the three main Eurasian Borrelia species, B. afzelii is adapted to rodents, whereas B. valaisiana and most strains of B. garinii are maintained by birds (12, 15, 16, 23, 26, 45). However, B. garinii OspA serotype 4 strains in Europe have been shown to be transmitted by rodents (17, 18) and, therefore, constitute a distinct ecotype within B. garinii. These strains have also been associated with high pathogenicity in humans, and their finer-scale geographical distribution seems highly focal (10, 34, 52, 53).In this study, we analyzed the intra- and interspecific phylogenetic relationships of B. burgdorferi, B. afzelii, B. garinii, B. valaisiana, B. lusitaniae, B. bissettii, and B. spielmanii by means of a novel MLSA scheme based on chromosomal housekeeping genes (30, 48).  相似文献   

10.
Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding to CD4 and a chemokine receptor, most commonly CCR5. CXCR4 is a frequent alternative coreceptor (CoR) in subtype B and D HIV-1 infection, but the importance of many other alternative CoRs remains elusive. We have analyzed HIV-1 envelope (Env) proteins from 66 individuals infected with the major subtypes of HIV-1 to determine if virus entry into highly permissive NP-2 cell lines expressing most known alternative CoRs differed by HIV-1 subtype. We also performed linear regression analysis to determine if virus entry via the major CoR CCR5 correlated with use of any alternative CoR and if this correlation differed by subtype. Virus pseudotyped with subtype B Env showed robust entry via CCR3 that was highly correlated with CCR5 entry efficiency. By contrast, viruses pseudotyped with subtype A and C Env proteins were able to use the recently described alternative CoR FPRL1 more efficiently than CCR3, and use of FPRL1 was correlated with CCR5 entry. Subtype D Env was unable to use either CCR3 or FPRL1 efficiently, a unique pattern of alternative CoR use. These results suggest that each subtype of circulating HIV-1 may be subject to somewhat different selective pressures for Env-mediated entry into target cells and suggest that CCR3 may be used as a surrogate CoR by subtype B while FPRL1 may be used as a surrogate CoR by subtypes A and C. These data may provide insight into development of resistance to CCR5-targeted entry inhibitors and alternative entry pathways for each HIV-1 subtype.Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding first to CD4 and then to a coreceptor (CoR), of which C-C chemokine receptor 5 (CCR5) is the most common (6, 53). CXCR4 is an additional CoR for up to 50% of subtype B and D HIV-1 isolates at very late stages of disease (4, 7, 28, 35). Many other seven-membrane-spanning G-protein-coupled receptors (GPCRs) have been identified as alternative CoRs when expressed on various target cell lines in vitro, including CCR1 (76, 79), CCR2b (24), CCR3 (3, 5, 17, 32, 60), CCR8 (18, 34, 38), GPR1 (27, 65), GPR15/BOB (22), CXCR5 (39), CXCR6/Bonzo/STRL33/TYMSTR (9, 22, 25, 45, 46), APJ (26), CMKLR1/ChemR23 (49, 62), FPLR1 (67, 68), RDC1 (66), and D6 (55). HIV-2 and simian immunodeficiency virus SIVmac isolates more frequently show expanded use of these alternative CoRs than HIV-1 isolates (12, 30, 51, 74), and evidence that alternative CoRs other than CXCR4 mediate infection of primary target cells by HIV-1 isolates is sparse (18, 30, 53, 81). Genetic deficiency in CCR5 expression is highly protective against HIV-1 transmission (21, 36), establishing CCR5 as the primary CoR. The importance of alternative CoRs other than CXCR4 has remained elusive despite many studies (1, 30, 70, 81). Expansion of CoR use from CCR5 to include CXCR4 is frequently associated with the ability to use additional alternative CoRs for viral entry (8, 16, 20, 63, 79) in most but not all studies (29, 33, 40, 77, 78). This finding suggests that the sequence changes in HIV-1 env required for use of CXCR4 as an additional or alternative CoR (14, 15, 31, 37, 41, 57) are likely to increase the potential to use other alternative CoRs.We have used the highly permissive NP-2/CD4 human glioma cell line developed by Soda et al. (69) to classify virus entry via the alternative CoRs CCR1, CCR3, CCR8, GPR1, CXCR6, APJ, CMKLR1/ChemR23, FPRL1, and CXCR4. Full-length molecular clones of 66 env genes from most prevalent HIV-1 subtypes were used to generate infectious virus pseudotypes expressing a luciferase reporter construct (19, 57). Two types of analysis were performed: the level of virus entry mediated by each alternative CoR and linear regression of entry mediated by CCR5 versus all other alternative CoRs. We thus were able to identify patterns of alternative CoR use that were subtype specific and to determine if use of any alternative CoR was correlated or independent of CCR5-mediated entry. The results obtained have implications for the evolution of env function, and the analyses revealed important differences between subtype B Env function and all other HIV-1 subtypes.  相似文献   

11.
Replication of hepatitis C virus (HCV) RNA occurs on intracellular membranes, and the replication complex (RC) contains viral RNA, nonstructural proteins, and cellular cofactors. We previously demonstrated that cyclophilin A (CyPA) is an essential cofactor for HCV infection and the intracellular target of cyclosporine''s anti-HCV effect. Here we investigate the mechanism by which CyPA facilitates HCV replication. Cyclosporine treatment specifically blocked the incorporation of NS5B into the RC without affecting either the total protein level or the membrane association of the protein. Other nonstructural proteins or viral RNAs in the RC were not affected. NS5B from the cyclosporine-resistant replicon was resistant to this disruption of RC incorporation. We also isolated membrane fractions from both naïve and HCV-positive cells and found that CyPA is recruited into membrane fractions in HCV-replicating cells via an interaction with RC-associated NS5B, which is sensitive to cyclosporine treatment. Finally, we introduced point mutations in the prolyl-peptidyl isomerase (PPIase) motif of CyPA and demonstrated a critical role of this motif in HCV replication in cDNA rescue experiments. We propose a model in which the incorporation of the HCV polymerase into the RC depends on its interaction with a cellular chaperone protein and in which cyclosporine inhibits HCV replication by blocking this critical interaction and the PPIase activity of CyPA. Our results provide a mechanism of action for the cyclosporine-mediated inhibition of HCV and identify a critical role of CyPA''s PPIase activity in the proper assembly and function of the HCV RC.Hepatitis C virus (HCV), of the family Flaviviridae, is an enveloped, positive-stranded RNA virus. Spread mostly by blood-borne transmission, HCV infects more than 170 million people worldwide. The viral genome is composed of a single open reading frame (ORF) plus 5′- and 3′-nontranslated regions. The ORF encodes a large polyprotein that is cleaved by cellular and viral proteases into 10 viral proteins. The structural proteins, including the capsid protein (core), two glycoproteins (E1 and E2), and a small ion channel protein (p7), reside in the N-terminal half of the polyprotein. The rest of the ORF encodes six nonstructural (NS) proteins: NS2, NS3, NS4A, NS4B, NS5A, and NS5B. NS3 through NS5B assemble into a replication complex (RC) and are necessary and sufficient for HCV RNA replication in cell culture (8, 42). NS3 is a multifunctional protein with both a serine protease and an RNA helicase activity. The protease activity is responsible for cleavage at the NS3-NS4A, NS4A-NS4B, NS4B-NS5A, and NS5A-NS5B junctions (5), and the helicase activity is probably required to unwind the double-stranded RNA intermediates formed during replication (38). NS4A serves as an essential cofactor for the NS3 protease and anchors the NS3 protein to intracellular membranes (25, 36, 39). NS4B induces the formation of a “membranous web” that is probably the site of HCV replication (16). It also contains a GTP-binding motif that is required for replication (17). The web is derived from the endoplasmic reticulum (ER) compartment, although proteins of early-endosome origin have also been found to locate to the web (62). NS5A is a phosphoprotein and an integral component of the viral RC. The precise function of NS5A in replication is still unknown but appears to be regulated by phosphorylation and its interaction with several cellular proteins (19, 22, 24, 51, 52, 59, 63, 67). In addition, it may be involved in the transition from replication and particle formation (4, 45, 64). NS5B is the RNA-dependent RNA polymerase that is responsible for copying the RNA genome of the virus during replication. Several cellular cofactors interact with NS5B and modulate its activity in the context of the viral RC (22, 24, 35, 69, 71).Positive-stranded RNA viruses alter the intracellular membranes of host cells to form an RC in which RNA replication occurs. Modifications include the proliferation and reorganization of certain cellular membranes (1). HCV forms an RC associated with altered cellular membranes (16, 23), and crude RCs (CRCs) that maintain the replicase activity in vitro can be isolated by membrane sedimentation or flotation techniques (2, 3, 18, 27, 37).Cyclosporine is a widely used immunosuppressive and anti-inflammatory drug for organ transplant patients. It functions by forming an inhibitory complex with cyclophilins (CyPs) that inhibits the phosphatase activity of calcineurin, which is important for T-cell activation. In recent years, cyclosporine and its derivatives have been shown to be highly effective in suppressing HCV replication in vitro (44, 49, 53, 68) and in vivo (30). The mechanism of this inhibition is independent of its immunosuppressive function and distinct from that of interferon (IFN) (44, 53, 56, 68).We recently showed that HCV infection in vitro is inhibited when CyPA, a major intracellular target of cyclosporine, is downregulated by RNA interference, and mutations in NS5B that confer cyclosporine-resistant binding to CyPA contribute to the cyclosporine resistance of the replicons harboring these mutations (56, 71). Here we report that CyPA is recruited into the HCV RC together with NS5B in HCV replicon or in HCV-infected cells. Cyclosporine disrupts the association between RC-incorporated NS5B and CyPA and results in an exclusion of the polymerase from the viral RC. We also show that the prolyl-peptidyl isomerase (PPIase) motif of CyPA is essential for HCV replication.  相似文献   

12.
The ability to undergo dramatic morphological changes in response to extrinsic cues is conserved in fungi. We have used the model yeast Schizosaccharomyces pombe to determine which intracellular signal regulates the dimorphic switch from the single-cell yeast form to the filamentous invasive growth form. The S. pombe Asp1 protein, a member of the conserved Vip1 1/3 inositol polyphosphate kinase family, is a key regulator of the morphological switch via the cAMP protein kinase A (PKA) pathway. Lack of a functional Asp1 kinase domain abolishes invasive growth which is monopolar, while an increase in Asp1-generated inositol pyrophosphates (PP) increases the cellular response. Remarkably, the Asp1 kinase activity encoded by the N-terminal part of the protein is regulated negatively by the C-terminal domain of Asp1, which has homology to acid histidine phosphatases. Thus, the fine tuning of the cellular response to environmental cues is modulated by the same protein. As the Saccharomyces cerevisiae Asp1 ortholog is also required for the dimorphic switch in this yeast, we propose that Vip1 family members have a general role in regulating fungal dimorphism.Eucaryotic cells are able to define and maintain a particular cellular organization and thus cellular morphology by executing programs modulated by internal and external signals. For example, signals generated within a cell are required for the selection of the growth zone after cytokinesis in the fission yeast Schizosaccharomyces pombe or the emergence of the bud in Saccharomyces cerevisiae (37, 44, 81). Cellular morphogenesis is also subject to regulation by a wide variety of external signals, such as growth factors, temperature, hormones, nutrient limitation, and cell-cell or cell-substrate contact (13, 34, 66, 75, 81). Both types of signals will lead to the selection of growth zones accompanied by the reorganization of the cytoskeleton.The ability to alter the growth form in response to environmental conditions is an important virulence-associated trait of pathogenic fungi which helps the pathogen to spread in and survive the host''s defense system (7, 32). Alteration of the growth form in response to extrinsic signals is not limited to pathogenic fungi but is also found in the model yeasts S. cerevisiae and S. pombe, in which it appears to represent a foraging response (1, 24).The regulation of polarized growth and the definition of growth zones have been studied extensively with the fission yeast S. pombe. In this cylindrically shaped organism, cell wall biosynthesis is restricted to one or both cell ends in a cell cycle-regulated manner and to the septum during cytokinesis (38). This mode of growth requires the actin cytoskeleton to direct growth and the microtubule cytoskeleton to define the growth sites (60). In interphase cells, microtubules are organized in antiparallel bundles that are aligned along the long axis of the cell and grow from their plus ends toward the cell tips. Upon contact with the cell end, microtubule growth will first pause and then undergo a catastrophic event and microtubule shrinkage (21). This dynamic behavior of the microtubule plus end is regulated by a disparate, conserved, microtubule plus end group of proteins, called the +TIPs. The +TIP complex containing the EB1 family member Mal3 is required for the delivery of the Tea1-Tea4 complex to the cell tip (6, 11, 27, 45, 77). The latter complex docks at the cell end and recruits proteins required for actin nucleation (46, 76). Thus, the intricate cross talk between the actin and the microtubule cytoskeleton at specific intracellular locations is necessary for cell cycle-dependent polarized growth of the fission yeast cell.The intense analysis of polarized growth control in single-celled S. pombe makes this yeast an attractive organism for the identification of key regulatory components of the dimorphic switch. S. pombe multicellular invasive growth has been observed for specific strains under specific conditions, such as nitrogen and ammonium limitation and the presence of excess iron (1, 19, 50, 61).Here, we have identified an evolutionarily conserved key regulator of the S. pombe dimorphic switch, the Asp1 protein. Asp1 belongs to the highly conserved family of Vip1 1/3 inositol polyphosphate kinases, which is one of two families that can generate inositol pyrophosphates (PP) (17, 23, 42, 54). The inositol polyphosphate kinase IP6K family, of which the S. cerevisiae Kcs1 protein is a member, is the “classical” family that can phosphorylate inositol hexakisphosphate (IP6) (70, 71). These enzymes generate a specific PP-IP5 (IP7), which has the pyrophosphate at position 5 of the inositol ring (20, 54). The Vip1 family kinase activity was unmasked in an S. cerevisiae strain with KCS1 and DDP1 deleted (54, 83). The latter gene encodes a nudix hydrolase (14, 68). The mammalian and S. cerevisiae Vip1 proteins phosphorylate the 1/3 position of the inositol ring, generating 1/3 diphosphoinositol pentakisphosphate (42). Both enzyme families collaborate to generate IP8 (17, 23, 42, 54, 57).Two modes of action have been described for the high-energy moiety containing inositol pyrophosphates. First, these molecules can phosphorylate proteins by a nonenzymatic transfer of a phosphate group to specific prephosphorylated serine residues (2, 8, 69). Second, inositol pyrophosphates can regulate protein function by reversible binding to the S. cerevisiae Pho80-Pho85-Pho81 complex (39, 40). This cyclin-cyclin-dependent kinase complex is inactivated by inositol pyrophosphates generated by Vip1 when cells are starved of inorganic phosphate (39, 41, 42).Regulation of phosphate metabolism in S. cerevisiae is one of the few roles specifically attributed to a Vip1 kinase. Further information about the cellular function of this family came from the identification of the S. pombe Vip1 family member Asp1 as a regulator of the actin nucleator Arp2/3 complex (22). The 106-kDa Asp1 cytoplasmic protein, which probably exists as a dimer in vivo, acts as a multicopy suppressor of arp3-c1 mutants (22). Loss of Asp1 results in abnormal cell morphology, defects in polarized growth, and aberrant cortical actin cytoskeleton organization (22).The Vip1 family proteins have a dual domain structure which consists of an N-terminal “rimK”/ATP-grasp superfamily domain found in certain inositol signaling kinases and a C-terminal part with homology to histidine acid phosphatases present in phytase enzymes (28, 53, 54). The N-terminal domain is required and sufficient for Vip1 family kinase activity, and an Asp1 variant with a mutation in a catalytic residue of the kinase domain is unable to suppress mutants of the Arp2/3 complex (17, 23, 54). To date, no function has been described for the C-terminal phosphatase domain, and this domain appears to be catalytically inactive (17, 23, 54).Here we describe a new and conserved role for Vip1 kinases in regulating the dimorphic switch in yeasts. Asp1 kinase activity is essential for cell-cell and cell-substrate adhesion and the ability of S. pombe cells to grow invasively. Interestingly, Asp1 kinase activity is counteracted by the putative phosphatase domain of this protein, a finding that allows us to describe for the first time a function for the C-terminal part of Vip1 proteins.  相似文献   

13.
Proper DNA methylation patterns are essential for mammalian development and differentiation. DNA methyltransferases (DNMTs) primarily establish and maintain global DNA methylation patterns; however, the molecular mechanisms for the generation and inheritance of methylation patterns are still poorly understood. We used sucrose density gradients of nucleosomes prepared by partial and maximum micrococcal nuclease digestion, coupled with Western blot analysis to probe for the interactions between DNMTs and native nucleosomes. This method allows for analysis of the in vivo interactions between the chromatin modification enzymes and their actual nucleosomal substrates in the native state. We show that little free DNA methyltransferase 3A and 3B (DNMT3A/3B) exist in the nucleus and that almost all of the cellular contents of DNMT3A/3B, but not DNMT1, are strongly anchored to a subset of nucleosomes. This binding of DNMT3A/3B does not require the presence of other well-known chromatin-modifying enzymes or proteins, such as proliferating cell nuclear antigen, heterochromatin protein 1, methyl-CpG binding protein 2, Enhancer of Zeste homolog 2, histone deacetylase 1, and UHRF1, but it does require an intact nucleosomal structure. We also show that nucleosomes containing methylated SINE and LINE elements and CpG islands are the main sites of DNMT3A/3B binding. These data suggest that inheritance of DNA methylation requires cues from the chromatin component in addition to hemimethylation.Proper DNA methylation patterns are essential for mammalian development and differentiation. More than three decades ago, de novo cytosine DNA methylation and its maintenance were proposed to exist in eukaryotic cells (29, 54); however, the molecular mechanisms for the generation and inheritance of methylation patterns are still poorly understood. DNA methyltransferases (DNMTs) DNMT1, DNMT3A, and DNMT3B primarily establish and maintain global DNA methylation patterns (39, 48). DNMT1 preferentially methylates hemimethylated DNA in vitro (7) and is tethered to replication foci during S phase (38). In contrast, DNMT3A and DNMT3B (DNMT3A/3B) have no preference for hemimethylated DNA (49) and are required for de novo methylation of genomic DNA (48). It has been thought that DNMT1 acts mainly as a “maintenance methyltransferase” during DNA synthesis and that DNMT3A and DNMT3B act as “de novo” enzymes. However, more recent studies indicate that DNMT1 may also be required for de novo methylation of genomic DNA (17, 30) and that DNMT3A/3B are also required for maintenance functions (11, 40, 55). Furthermore, the different DNMTs cooperate in maintaining the methylation of some regions of the genome, particularly repetitive elements (40, 53).Recruitment of individual DNMT enzymes to different regions of chromatin in vivo, particularly to gene regulatory regions, may require interaction with auxiliary factors (28, 36). DNMT1, which is diffusely localized throughout nuclei in non-S-phase cells (38), is targeted to replication foci by interacting with proliferating cell nuclear antigen (PCNA) (15) and also physically interacts with UHRF1 (ubiquitinlike, containing PHD and RING finger domains 1) that binds to hemimethylated DNA (3, 4, 8, 27, 62). DNMT3 enzymes are usually found localized to heterochromatin regions in most transient-expression assays (5, 12). As genomic DNA in chromatin is packaged into nucleosomes which might limit the accessibility of target sites to the enzymes, the interaction of DNMTs with nucleosomes in a chromatin context is important for the regulation of genomic methylation.Genetic and biochemical studies have provided many insights into the distinct and cooperative functions of the DNMT enzymes; however, few of these studies have addressed how they interact with chromatin in vivo. Recombinant DNMT1 and DNMT3 enzymes can methylate the CpG sites on nucleosomes assembled in vitro (26, 50, 56, 65). Recently DNMT3L has been found to connect DNMT3A2 to nucleosomes in embryonic stem cells (52). However, DNMT3L is expressed only during gametogenesis and embryonic stages (1, 9), suggesting that other mechanisms might be necessary for directing the enzyme to specific chromatin regions in somatic cells.In the present study, we investigated how different DNMT enzymes interact with chromatin at the nucleosomal level in somatic cell lines. Micrococcal nuclease (MNase) treatment of nuclei in a low-ionic-strength buffer digests nucleosomal linker DNA regions, thereby minimizing the disruption of protein complexes on the nucleosomes. We prepared nucleosomes from partial or maximum MNase-digested nuclei and resolved them on sucrose density gradients to analyze their interactions with chromatin proteins. The results indicate that while DNMT1 interacts primarily with linker DNA, DNMT3A/3B enzymes interact strongly with nucleosomes containing methylated repetitive elements and also containing methylated CpG islands (CGIs) and may not require additional proteins for this strong binding. These data are particularly intriguing in that they provide insights into the mechanisms of the interaction of DNMTs with chromatin and maintenance of DNA methylation in somatic cells.  相似文献   

14.
Human cytomegalovirus (HCMV) UL37 proteins traffic sequentially from the endoplasmic reticulum (ER) to the mitochondria. In transiently transfected cells, UL37 proteins traffic into the mitochondrion-associated membranes (MAM), the site of contact between the ER and mitochondria. In HCMV-infected cells, the predominant UL37 exon 1 protein, pUL37x1, trafficked into the ER, the MAM, and the mitochondria. Surprisingly, a component of the MAM calcium signaling junction complex, cytosolic Grp75, was increasingly enriched in heavy MAM from HCMV-infected cells. These studies show the first documented case of a herpesvirus protein, HCMV pUL37x1, trafficking into the MAM during permissive infection and HCMV-induced alteration of the MAM protein composition.The human cytomegalovirus (HCMV) UL37 immediate early (IE) locus expresses multiple products, including the predominant UL37 exon 1 protein, pUL37x1, also known as viral mitochondrion-localized inhibitor of apoptosis (vMIA), during lytic infection (16, 22, 24, 39, 44). The UL37 glycoprotein (gpUL37) shares UL37x1 sequences and is internally cleaved, generating pUL37NH2 and gpUL37COOH (2, 22, 25, 26). pUL37x1 is essential for the growth of HCMV in humans (17) and for the growth of primary HCMV strains (20) and strain AD169 (14, 35, 39, 49) but not strain TownevarATCC in permissive human fibroblasts (HFFs) (27).pUL37x1 induces calcium (Ca2+) efflux from the endoplasmic reticulum (ER) (39), regulates viral early gene expression (5, 10), disrupts F-actin (34, 39), recruits and inactivates Bax at the mitochondrial outer membrane (MOM) (4, 31-33), and inhibits mitochondrial serine protease at late times of infection (28).Intriguingly, HCMV UL37 proteins localize dually in the ER and in the mitochondria (2, 9, 16, 17, 24-26). In contrast to other characterized, similarly localized proteins (3, 6, 11, 23, 30, 38), dual-trafficking UL37 proteins are noncompetitive and sequential, as an uncleaved gpUL37 mutant protein is ER translocated, N-glycosylated, and then imported into the mitochondria (24, 26).Ninety-nine percent of ∼1,000 mitochondrial proteins are synthesized in the cytosol and directly imported into the mitochondria (13). However, the mitochondrial import of ER-synthesized proteins is poorly understood. One potential pathway is the use of the mitochondrion-associated membrane (MAM) as a transfer waypoint. The MAM is a specialized ER subdomain enriched in lipid-synthetic enzymes, lipid-associated proteins, such as sigma-1 receptor, and chaperones (18, 45). The MAM, the site of contact between the ER and the mitochondria, permits the translocation of membrane-bound lipids, including ceramide, between the two organelles (40). The MAM also provides enriched Ca2+ microdomains for mitochondrial signaling (15, 36, 37, 43, 48). One macromolecular MAM complex involved in efficient ER-to-mitochondrion Ca2+ transfer is comprised of ER-bound inositol 1,4,5-triphosphate receptor 3 (IP3R3), cytosolic Grp75, and a MOM-localized voltage-dependent anion channel (VDAC) (42). Another MAM-stabilizing protein complex utilizes mitofusin 2 (Mfn2) to tether ER and mitochondrial organelles together (12).HCMV UL37 proteins traffic into the MAM of transiently transfected HFFs and HeLa cells, directed by their NH2-terminal leaders (8, 47). To determine whether the MAM is targeted by UL37 proteins during infection, we fractionated HCMV-infected cells and examined pUL37x1 trafficking in microsomes, mitochondria, and the MAM throughout all temporal phases of infection. Because MAM domains physically bridge two organelles, multiple markers were employed to verify the purity and identity of the fractions (7, 8, 19, 46, 47).(These studies were performed in part by Chad Williamson in partial fulfillment of his doctoral studies in the Biochemistry and Molecular Genetics Program at George Washington Institute of Biomedical Sciences.)HFFs and life-extended (LE)-HFFs were grown and not infected or infected with HCMV (strain AD169) at a multiplicity of 3 PFU/cell as previously described (8, 26, 47). Heavy (6,300 × g) and light (100,000 × g) MAM fractions, mitochondria, and microsomes were isolated at various times of infection and quantified as described previously (7, 8, 47). Ten- or 20-μg amounts of total lysate or of subcellular fractions were resolved by SDS-PAGE in 4 to 12% Bis-Tris NuPage gels (Invitrogen) and examined by Western analyses (7, 8, 26). Twenty-microgram amounts of the fractions were not treated or treated with proteinase K (3 μg) for 20 min on ice, resolved by SDS-PAGE, and probed by Western analysis. The blots were probed with rabbit anti-UL37x1 antiserum (DC35), goat anti-dolichyl phosphate mannose synthase 1 (DPM1), goat anti-COX2 (both from Santa Cruz Biotechnology), mouse anti-Grp75 (StressGen Biotechnologies), and the corresponding horseradish peroxidase-conjugated secondary antibodies (8, 47). Reactive proteins were detected by enhanced chemiluminescence (ECL) reagents (Pierce), and images were digitized as described previously (26, 47).  相似文献   

15.
The Oxa1 protein plays a central role in facilitating the cotranslational insertion of the nascent polypeptide chains into the mitochondrial inner membrane. Mitochondrially encoded proteins are synthesized on matrix-localized ribosomes which are tethered to the inner membrane and in physical association with the Oxa1 protein. In the present study we used a chemical cross-linking approach to map the Saccharomyces cerevisiae Oxa1-ribosome interface, and we demonstrate here a close association of Oxa1 and the large ribosomal subunit protein, MrpL40. Evidence to indicate that a close physical and functional relationship exists between MrpL40 and another large ribosomal protein, the Mrp20/L23 protein, is also provided. MrpL40 shares sequence features with the bacterial ribosomal protein L24, which like Mrp20/L23 is known to be located adjacent to the ribosomal polypeptide exit site. We propose therefore that MrpL40 represents the Saccharomyces cerevisiae L24 homolog. MrpL40, like many mitochondrial ribosomal proteins, contains a C-terminal extension region that bears no similarity to the bacterial counterpart. We show that this C-terminal mitochondria-specific region is important for MrpL40''s ability to support the synthesis of the correct complement of mitochondrially encoded proteins and their subsequent assembly into oxidative phosphorylation complexes.The mitochondrial genome encodes a small, but important, number of proteins (8). These proteins are predominantly essential components of the mitochondrial oxidative phosphorylation (OXPHOS) machinery. In the yeast Saccharomyces cerevisiae the proteins encoded by the mitochondrial DNA (mtDNA) include cytochrome c oxidase subunits Cox1, Cox2, and Cox3, cytochrome b of the cytochrome bc1 complex, F1Fo-ATP synthase subunits Atp6, Atp8, and Atp9, and the small ribosomal subunit component Var1. With the exception of Var1, these mitochondrially encoded proteins are integral membrane proteins which become inserted into the inner membrane during their synthesis on mitochondrial ribosomes tethered to the inner membrane (11, 19, 29, 32, 34). The cotranslational membrane insertion of these proteins is achieved by maintaining a close physical association of the ribosomes to the inner membrane at sites where the insertion machinery exists (19, 31, 32).Oxa1 is an inner membrane protein that forms a central component of the insertion machinery, whose presence is required for the cotranslational membrane insertion of the mitochondrially encoded proteins (4-6, 15-17). The Oxa1 protein has been shown to physically associate with the ribosomes and more specifically with the large ribosomal subunit. Matrix-exposed elements of the Oxa1 protein, such as its hydrophilic C-terminal tail, support this Oxa1-ribosome interaction (19, 32). Furthermore, in intact mitochondria we have previously demonstrated that Oxa1 can be chemically cross-linked to Mrp20, a component of the large ribosomal subunit (19). Mrp20 is homologous to the bacterial ribosomal protein L23, a component known from the structural analysis of the ribosomes to be located next to the polypeptide exit site of the large ribosomal subunit (3, 10, 23, 27, 30). Thus, it was concluded that Oxa1, the site of membrane insertion into the inner membrane, exists in close physical proximity to the large ribosomal subunit and specifically to that region of the ribosomes where the nascent chain emerges. This close physical relationship between ribosomal components and the Oxa1 insertion site has been proposed to support a tight coordination between the protein translation and membrane insertion events (19, 31, 32). Given the strong hydrophobicity of the OXPHOS complex subunits which are encoded by the mitochondrial DNA and synthesized by these ribosomes, a close coupling of the translation and insertion events is proposed to ensure that the hydrophobic nascent chains are directly inserted into the membrane during their synthesis. The exposure of hydrophobic nascent chains to the hydrophilic matrix space may promote their aggregation and thus incompetency for subsequence membrane insertion.In bacteria, the L23 protein has been implicated to play a direct role in the cotranslational insertion of proteins into the membrane (7, 13, 24, 33). Thus, it is possible that proteins adjacent to the polypeptide exit site of mitochondrial ribosomes may be directly involved in targeting ribosomes to specific regions of the inner membrane where the membrane insertion and subsequent assembly events occur. The mitochondrial ribosomes resemble their prokaryotic ancestors in some respects, e.g., antibiotic sensitivity, but they differ in a number of important ways (1, 12, 22, 30). In general, the protein content of the mitochondrial ribosomes is greater than their bacterial counterparts. This increase in protein content is largely attributed to the fact that the mitochondrial ribosomal proteins are larger in size than their bacterial homologs. Over the course of evolution, many of the mitochondrial ribosomal proteins have acquired novel extensions, new domains, in addition to their bacterial homology domains. These acquired extensions not only include N-terminal (often cleavable) signals to target these proteins (nuclear encoded) to the mitochondria but also in many instances large C-terminal extensions, which are unique to the mitochondrial ribosomal proteins and have thus been termed “mitospecific domains” (12, 30). Largely uncharacterized, the functional relevance of these various mitospecific domains of the ribosomal proteins remains unknown. It is speculated that some (or all) of these mitospecific domains serve to ensure that the ribosome becomes assembled and is translationally active while bound to the inner membrane surface.In the present study we sought to further characterize the interaction of the mitochondrial ribosome with the Oxa1 protein. We show here that MrpL40, a large ribosomal subunit component, is physically close to both the Mrp20 and Oxa1 proteins, demonstrating the proximity of MrpL40 to both the ribosomal polypeptide exit site and the Oxa1 membrane insertion site. MrpL40 contains a large C-terminal mitospecific domain, which includes a predicted α-helical region at its extreme C-terminal end. The results presented here highlight that the integrity of this domain of MrpL40 is crucial to ensure ribosome translational fidelity and subsequent OXPHOS complex assembly.  相似文献   

16.
17.
18.
19.
20.
Deformylases are metalloproteases in bacteria, plants, and humans that remove the N-formyl-methionine off peptides in vitro. The human homolog of peptide deformylase (HsPDF) resides in the mitochondria, along with its putative formylated substrates; however, the cellular function of HsPDF remains elusive. Here we report on the function of HsPDF in mitochondrial translation and oxidative phosphorylation complex biogenesis. Functional HsPDF appears to be necessary for the accumulation of mitochondrial DNA-encoded proteins and assembly of new respiratory complexes containing these proteins. Consequently, inhibition of HsPDF reduces respiratory function and cellular ATP levels, causing dependence on aerobic glycolysis for cell survival. A series of structurally different HsPDF inhibitors and control peptidase inhibitors confirmed that inhibition of HsPDF decreases mtDNA-encoded protein accumulation. Therefore, HsPDF appears to have a role in maintenance of mitochondrial respiratory function, and this function is analogous to that of chloroplast PDF.The human mitochondrial protein peptide deformylase, HsPDF, is a metalloprotease that removes the formyl moiety on the methionine of N-formyl-methionine peptide substrates in an enzymatic assay (24, 35). Despite the slow kinetic properties of HsPDF in an in vitro deformylation assay (24, 29, 35), we have shown that small interfering RNA (siRNA) interference of HsPDF decreases human cancer cell proliferation. Similarly, pharmacologic inhibition with the PDF antibiotic inhibitor actinonin and its analogs results in mitochondrial membrane depolarization and promotes cell death or proliferation arrest in a wide variety of cancer cell lines (18, 25). However, the cellular function of HsPDF remains elusive, and others have proposed that it has none (29). In bacteria, deformylation of nascent peptides is necessary for removal of the N-terminal methionine (36) and posttranslational processing of at least a subset of proteins that contribute to cell growth and viability (28). Prokaryotic PDF thus fulfills a role in cotranslational processing (7) and in protein degradation (41).In mammals, N-terminal formylation of proteins is only known to occur during mitochondrial translation initiation, as in prokaryotic protein translation (6). In contrast to bacteria, where the entire proteome is formylated for translation initiation, formylation in eukaryotes is limited to the 13 mitochondrial DNA (mtDNA)-encoded proteins. Formylation is important for mitochondrial translation, because formyl-Met-tRNA, but not Met-tRNA, is recognized by initiation factor 2 as the initiator tRNA (26, 37, 39). Therefore, the participation of HsPDF in protein post- or cotranslational processing can be narrowed down to these mitochondrial translation products.Despite the current understanding of the function of formyl-methionine in the initiation of protein synthesis in mammalian mitochondria (38, 39), the functional relevance of the downstream processing of nascent mitochondrial translation products has remained unexplored. Furthermore, it has been assumed that human mitochondria-encoded proteins, like those of bovine origin, are generally not deformylated after synthesis (45).The mammalian mitochondrial genome-encoded proteins are all subunits of four of the five oxidative phosphorylation respiratory chain enzyme complexes (I, III, IV, and V) (2, 40, 42). Respiratory complexes are comprised of multiple proteins. With the exception of complex II, which is comprised entirely of nuclear DNA-encoded subunits, all other complexes include both nuclear and mitochondrial DNA-encoded proteins. Synthesis of key mtDNA-encoded protein subunits, and the assembly of these proteins with multiple nuclear-encoded subunits within the mitochondria, is necessary for the function of each individual complex (16, 30, 44). Moreover, a functional interdependence among stably assembled respiratory complexes has been demonstrated (1). Mutations in human mtDNA that affect protein-coding regions or nuclear DNA mutations that affect expression of respiratory complex subunits cause disease (13), including Parkinson''s disease, for example, in which decreased respiratory function and compromised cell viability have been demonstrated (5, 21, 23). Therefore, the importance of properly assembled mitochondrial respiratory complexes suggests that their disruption, by inhibition of mtDNA-encoded protein processing, could have significant effects on cellular function.We hypothesized that HsPDF-mediated processing of mtDNA-encoded proteins is necessary for proper function of the respiratory chain complexes. To determine how the human deformylase activity contributes to cellular function, we used pharmacologic inhibition of HsPDF activity with the hydroxamic acid peptidomimetic inhibitor of PDF, actinonin, and confirmed our findings with a variety of other structurally different inhibitors. PDF has been shown to be a target of actinonin in bacteria (9), human cells (24), and plants (17).Here we show that inhibition of HsPDF function in mitochondria of human cell lines reduces mtDNA-encoded protein accumulation, new respiratory complex assembly, and energy production by the mitochondria. Aerobic glycolysis-dependent cell survival ensues upon disruption of HsPDF function. Therefore, HsPDF appears to fulfill a function in the mitochondria and to have a role in mtDNA-encoded protein-containing oxidative phosphorylation (OXPHOS) complex biogenesis.  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