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 828 毫秒
1.
The budding yeast formins, Bnr1 and Bni1, behave very differently with respect to their interactions with muscle actin. However, the mechanisms underlying these differences are unclear, and these formins do not interact with muscle actin in vivo. We use yeast wild type and mutant actins to further assess these differences between Bnr1 and Bni1. Low ionic strength G-buffer does not promote actin polymerization. However, Bnr1, but not Bni1, causes the polymerization of pyrene-labeled Mg-G-actin in G-buffer into single filaments based on fluorometric and EM observations. Polymerization by Bnr1 does not occur with Ca-G-actin. By cosedimentation, maximum filament formation occurs at a Bnr1:actin ratio of 1:2. The interaction of Bnr1 with pyrene-labeled S265C Mg-actin yields a pyrene excimer peak, from the cross-strand interaction of pyrene probes, which only occurs in the context of F-actin. In F-buffer, Bnr1 promotes much faster yeast actin polymerization than Bni1. It also bundles the F-actin in contrast to the low ionic strength situation where only single filaments form. Thus, the differences previously observed with muscle actin are not actin isoform-specific. The binding of both formins to F-actin saturate at an equimolar ratio, but only about 30% of each formin cosediments with F-actin. Finally, addition of Bnr1 but not Bni1 to pyrene-labeled wild type and S265C Mg-F actins enhanced the pyrene- and pyrene-excimer fluorescence, respectively, suggesting Bnr1 also alters F-actin structure. These differences may facilitate the ability of Bnr1 to form the actin cables needed for polarized delivery of nutrients and organelles to the growing yeast bud.Bni1 and Bnr1 are the two formin isoforms expressed in Saccharomyces cerevisiae (1, 2). These proteins, as other isoforms in the formin family, are large multidomain proteins (3, 4). Several regulatory domains, including one for binding the G-protein rho, are located at the N-terminal half of the protein (47). FH1, FH2, and Bud6 binding domains are located in the C-terminal half of the protein (8). The formin homology 1 (FH1)2 domain contains several sequential poly-l-proline motifs, and it interacts with the profilin/actin complex to recruit actin monomers and regulate the insertion of actin monomers at the barbed end of actin (911). The fomin homology domain 2 (FH2) forms a donut-shaped homodimer, which wraps around actin dimers at the barbed end of actin filaments (12, 13). One important function of formin is to facilitate actin polymerization by stabilizing actin dimers or trimers under polymerization conditions and then to processively associate with the barbed end of the elongating filament to control actin filament elongation kinetics (1318).A major unsolved protein in the study of formins is the elucidation of the individual functions of different isoforms and their regulation. In vivo, these two budding yeast formins have distinct cellular locations and dynamics (1, 2, 19, 20). Bni1 concentrates at the budding site before the daughter cell buds from the mother cell, moves along with the tip of the daughter cell, and then travels back to the neck between daughter and mother cells at the end of segregation. Bnr1 localizes only at the neck of the budding cell in a very short period of time after bud emergence. Although a key cellular function of these two formins in yeast is to promote actin cable formation (8, 18), the roles of the individual formins in different cellular process is unclear because deleting either individual formin gene has limited impact on cell growth and deleting both genes together is lethal (21).Although each of the two formins can nucleate actin filament formation in vitro, the manner in which they affect polymerization is distinctly isoform-specific. Most of this mechanistic work in vitro has used formin fragments containing the FH1 and FH2 domains. Bni1 alone processively caps the barbed end of actin filaments partially inhibiting polymerization at this end (14, 16, 18). The profilin-actin complex, recruited to the actin barbed end through its binding to Bni1 FH1 domain, possibly raises the local actin concentration and appears to allow this inhibition to be overcome, thereby, accelerating barbed end polymerization. It has also been shown that this complex modifies the kinetics of actin dynamics at the barbed end (9, 11, 18, 22). Moreover, Bni1 participation leads only to the formation of single filaments (8). In comparison, the Bnr1 FH1-FH2 domain facilitates actin polymerization much more efficiently than does Bni1. Moseley and Goode (8) showed Bnr1 accelerates actin polymerization up to 10 times better than does Bni and produces actin filament bundles when the Bnr1/actin molar ratio is above 1:2. Finally, the regulation of Bni1 and Bnr1 by formin binding is different. For example, Bud 6/Aip3, a yeast cell polarity factor, binds to Bni1, but not Bnr1, and also stimulates its activity in vitro.For their studies, Moseley and Goode (8) utilized mammalian skeletal muscle actin instead of the S. cerevisiae actin with which the yeast formins are designed to function. It is entirely possible that the differences observed with the two formins are influenced quantitatively or qualitatively by the nature of the actin used in the study. This possibility must be seriously considered because although yeast and muscle actins are 87% identical in sequence, they display marked differences in their polymerization behavior (23). Yeast actin nucleates filaments better than muscle actin (24, 25). It appears to form shorter and more flexible filaments than does muscle actin (26, 27). Finally, the disposition of the Pi released during the hydrolysis of ATP that occurs during polymerization is different. Yeast actin releases its Pi concomitant with hydrolysis of the bound ATP whereas muscle actin retains the Pi for a significant amount of time following nucleotide hydrolysis (28, 29). This difference is significant because ADP-Pi F-actin has been shown to be more stable than ADP F-actin (30). Another example of this isoform dependence is the interaction of yeast Arp2/3 with yeast versus muscle actins (31). Yeast Arp2/3 complex accelerates polymerization of muscle actin only in the presence of a nucleation protein factor such as WASP. However, with yeast actin, no such auxiliary protein is required. In light of these actin behavioral differences, to better understand the functional differences of these two formins in vivo, we have studied the behavior of Bni 1 and Bnr 1 with WT and mutant yeast actins, and we have also explored the molecular basis underlying the Bnr 1-induced formation of actin nuclei from G-actin.  相似文献   

2.
Vinculin links integrins to the actin cytoskeleton by binding F-actin. Little is known with respect to how this interaction occurs or affects actin dynamics. Here we assess the consequence of the vinculin tail (VT) on actin dynamics by examining its binding to monomeric and filamentous yeast actins. VT causes pyrene-labeled G-actin to polymerize in low ionic strength buffer (G-buffer), conditions that normally do not promote actin polymerization. Analysis by electron microscopy shows that, under these conditions, the filaments form small bundles at low VT concentrations, which gradually increase in size until saturation occurs at a ratio of 2 VT:1 actin. Addition of VT to pyrene-labeled mutant yeast G-actin (S265C) produced a fluorescence excimer band, which requires a relatively normal filament geometry. In higher ionic strength polymerization-promoting F-buffer, substoichiometric amounts of VT accelerate the polymerization of pyrene-labeled WT actin. However, the amplitude of the pyrene fluorescence caused by actin polymerization is quenched as the VT concentration increases without an effect on net actin polymerization as determined by centrifugation assays. Finally, addition of VT to preformed pyrene-labeled S265C F-actin causes a concentration-dependent decrease in the maximum amplitude of the pyrene fluorescence band demonstrating the ability of VT to remodel the conformation of the actin filament. These observations support the idea that vinculin can link adhesion plaques to the cytoskeleton by initiating the formation of bundled actin filaments or by remodeling existing filaments.Cell migration is critical for embryonic development, adult homeostasis, inflammatory responses, and wound healing. To migrate, a cell must coordinate a number of different inputs into appropriate cellular responses. The cell must polarize in the direction of migration and extend lamellipodial and/or filopodial protrusions. Nascent adhesions that assemble within the branched actin network of the lamellipodium must link to the underlying actin cytoskeleton. This process allows for the maturation of adhesions to structures that anchor the protrusion. These adhesions also provide the traction forces necessary to pull the cell body forward and break older adhesions at the cell rear. Perturbation of any of these events affects a cell''s migratory ability. For example, nascent adhesions that do not form linkages to the actin cytoskeleton cannot effectively anchor the protrusion to the substratum. The result is an extension that folds back upon itself, forming a membrane ruffle that cannot provide the traction forces necessary for migration.How adhesions establish links to the underlying actin cytoskeleton has been an area of intense investigation. Integrin-containing structures are active areas of actin polymerization suggesting that adhesion plaques can initiate actin filament formation (reviewed in Refs. 13). Focal complexes are small integrin clusters that are found exclusively at the tips of lamellipodia and filopodia. Formation of these structures is closely coupled with actin assembly in protruding regions of cells. Accumulating evidence indicates that adhesion complex components recruit the Arp2/3 complex, a potent nucleator of actin polymerization. Our work (4) and that of others (57) demonstrates that the Arp2/3 complex is recruited to focal complexes or transient adhesion structures reminiscent of focal complexes by binding vinculin. FAK has also been implicated in linking focal complexes to the actin cytoskeleton by virtue of its ability to recruit and activate the Arp2/3 complex (8). Furthermore, efficient focal complex assembly requires the actin-binding protein, cortactin, which could affect adhesion assembly by interacting with the Arp2/3 complex (9). Hence, many of the known mechanisms for initiating filament formation involve recruitment of the Arp2/3 complex, which initiates the formation of branched actin filaments (55). It is surprising then that the earliest detectable forms of actin-associated adhesions are interconnected by short actin bundles, not branched filaments (10). These observations suggest that our current understanding for how nascent adhesions initiate filament formation is incomplete.The earliest detectable actin-associated adhesions are “dots or doublets of dots” and are highly enriched in integrins, paxillin, and vinculin (10), suggesting that one of these molecules has the capability to initiate actin filament formation from such a plaque. Vinculin has long been implicated in linking adhesion plaques to the actin cytoskeleton by virtue of the ability of its tail to bind (11) and bundle F-actin (12). The interaction of vinculin with actin has been extensively studied from the perspective of vinculin (11, 1323). Studies of recombinant proteins identified two regions of the vinculin tail (VT)2 that bind F-actin independently (21, 17), but mapping these sites onto the VT crystal structure reveals that these peptides do not correspond to distinct sites (25). Upon binding actin, vinculin undergoes a conformational change that promotes dimerization suggesting that vinculin self-association may be important for its bundling activities (15).Less is known with respect to the effect of vinculin on actin filament formation and structure. This lack of knowledge stems from the fact that many of the early studies showed vinculin to have no effect on actin dynamics (2628). However, these experiments were performed using chicken gizzard vinculin, which exists almost exclusively in a conformation where the actin binding sites are inaccessible, or from preparations that contain contaminants that produce false negatives (29). More recently, recombinant VT proteins were shown to cross-link and bundle actin (23). However, the interaction of vinculin with G-actin and the effect of vinculin on actin filament dynamics have not been explored. In this study, we have assessed the interaction of vinculin with pyrene-labeled wild-type and mutant yeast actins. We show that the VT can promote the formation of an actin nucleus from which filaments arise and alter the assembly and structure of actin filaments. These findings provide novel insights into how adhesion plaques may be linked to the actin cytoskeleton.  相似文献   

3.
This study demonstrates the utility of Lifeact for the investigation of actin dynamics in Neurospora crassa and also represents the first report of simultaneous live-cell imaging of the actin and microtubule cytoskeletons in filamentous fungi. Lifeact is a 17-amino-acid peptide derived from the nonessential Saccharomyces cerevisiae actin-binding protein Abp140p. Fused to green fluorescent protein (GFP) or red fluorescent protein (TagRFP), Lifeact allowed live-cell imaging of actin patches, cables, and rings in N. crassa without interfering with cellular functions. Actin cables and patches localized to sites of active growth during the establishment and maintenance of cell polarity in germ tubes and conidial anastomosis tubes (CATs). Recurrent phases of formation and retrograde movement of complex arrays of actin cables were observed at growing tips of germ tubes and CATs. Two populations of actin patches exhibiting slow and fast movement were distinguished, and rapid (1.2 μm/s) saltatory transport of patches along cables was observed. Actin cables accumulated and subsequently condensed into actin rings associated with septum formation. F-actin organization was markedly different in the tip regions of mature hyphae and in germ tubes. Only mature hyphae displayed a subapical collar of actin patches and a concentration of F-actin within the core of the Spitzenkörper. Coexpression of Lifeact-TagRFP and β-tubulin–GFP revealed distinct but interrelated localization patterns of F-actin and microtubules during the initiation and maintenance of tip growth.Actins are highly conserved proteins found in all eukaryotes and have an enormous variety of cellular roles. The monomeric form (globular actin, or G-actin) can self-assemble, with the aid of numerous actin-binding proteins (ABPs), into microfilaments (filamentous actin, or F-actin), which, together with microtubules, form the two major components of the fungal cytoskeleton. Numerous pharmacological and genetic studies of fungi have demonstrated crucial roles for F-actin in cell polarity, exocytosis, endocytosis, cytokinesis, and organelle movement (6, 7, 20, 34, 35, 51, 52, 59). Phalloidin staining, immunofluorescent labeling, and fluorescent-protein (FP)-based live-cell imaging have revealed three distinct subpopulations of F-actin-containing structures in fungi: patches, cables, and rings (1, 14, 28, 34, 60, 63, 64). Actin patches are associated with the plasma membrane and represent an accumulation of F-actin around endocytic vesicles (3, 26, 57). Actin cables are bundles of actin filaments stabilized with cross-linking proteins, such as tropomyosins and fimbrin, and are assembled by formins at sites of active growth, where they form tracks for myosin V-dependent polarized secretion and organelle transport (10, 16, 17, 27, 38, 47, 48). Cables, unlike patches, are absolutely required for polarized growth in the budding yeast Saccharomyces cerevisiae (34, 38). Contractile actomyosin rings are essential for cytokinesis in budding yeast, whereas in filamentous fungi, actin rings are less well studied but are known to be involved in septum formation (20, 28, 34, 39, 40).Actin cables and patches have been particularly well studied in budding yeast. However, there are likely to be important differences between F-actin architecture and dynamics in budding yeast and those in filamentous fungi, as budding yeasts display only a short period of polarized growth during bud formation, which is followed by isotropic growth over the bud surface (10). Sustained polarized growth during hyphal morphogenesis is a defining feature of filamentous fungi (21), making them attractive models for studying the roles of the actin cytoskeleton in cell polarization, tip growth, and organelle transport.In Neurospora crassa and other filamentous fungi, disruption of the actin cytoskeleton leads to rapid tip swelling, which indicates perturbation of polarized tip growth, demonstrating a critical role for F-actin in targeted secretion to particular sites on the plasma membrane (7, 22, 29, 56). Immunofluorescence studies of N. crassa have shown that F-actin localizes to hyphal tips as “clouds” and “plaques” (7, 54, 59). However, immunolabeling has failed to reveal actin cables in N. crassa and offers limited insights into F-actin dynamics. Live-cell imaging of F-actin architecture and dynamics has not been accomplished in N. crassa, yet it is expected to yield key insights into cell polarization, tip growth, and intracellular transport.We took advantage of a recently developed live-cell imaging probe for F-actin called Lifeact (43). Lifeact is a 17-amino-acid peptide derived from the N terminus of the budding yeast actin-binding protein Abp140 (5, 63) and has recently been demonstrated to be a universal live-cell imaging marker for F-actin in eukaryotes (43). Here, we report the successful application of fluorescent Lifeact fusion constructs for live-cell imaging of F-actin in N. crassa. We constructed two synthetic genes consisting of Lifeact fused to “synthetic” green fluorescent protein (sGFP) (S65T) (henceforth termed GFP) (12) or red fluorescent protein (TagRFP) (33) and expressed these constructs in various N. crassa strains. In all strain backgrounds, fluorescent Lifeact constructs clearly labeled actin patches, cables, and rings and revealed a direct association of F-actin structures with sites of cell polarization and active tip growth. Our results demonstrate the efficacy of Lifeact as a nontoxic live-cell imaging probe in N. crassa.  相似文献   

4.
Auditory hair cell function requires proper assembly and regulation of the nonmuscle gamma isoactin-rich cytoskeleton, and six point mutations in this isoactin cause a type of delayed onset autosomal dominant nonsyndromic progressive hearing loss, DFNA20/26. The molecular basis underlying this actin-dependent hearing loss is unknown. To address this problem, the mutations have been introduced into yeast actin, and their effects on actin function were assessed in vivo and in vitro. Because we previously showed that polymerization was unaffected in five of the six mutants, we have focused on proteins that regulate actin, in particular cofilin, which severs F-actin and sequesters actin monomers. The mutations do not affect the interaction of cofilin with G-actin. However, T89I and V370A mutant F-actins are much more susceptible to cofilin disassembly than WT filaments in vitro. Conversely, P332A filaments demonstrate enhanced resistance. Wild type actin solutions containing T89I, K118M, or P332A mutant actins at mole fractions similar to those found in the hair cell respond in vitro toward cofilin in a manner proportional to the level of the mutant present. Finally, depression of cofilin action in vivo by elimination of the cofilin-activating protein, Aip1p, rescues the inability to grow on glycerol caused by K118M, T278I, P332A, and V370A. These results suggest that a filament instability caused by these mutations can be balanced by decreasing a system in vivo that promotes increased filament turnover. Such mutant-dependent filament destabilization could easily result in hair cell malfunction leading to the late-onset hearing loss observed in these patients.Our ability to perceive sounds hinges on the proper functioning of a highly specialized group of cells, called hair cells, which are housed in the cochlea of the inner ear. Their actin cytoskeleton plays a vital role in the ability of these cells to transduce mechanical stimuli into electrical responses (1). The columnar auditory hair cell consists of three actin-rich components (2). First are the three graded rows of finger-like projections from the apical surface of the cell referred to as stereocilia. These protrusions, comprised predominantly of bundles of actin filaments, are surrounded by the cellular membrane (3, 4). Sound-generated pressure waves in the cochlear fluid displace the basilar membrane, which in turn forces the stereocilia into the overlaying tectorial membrane, resulting in their mechanical deformation. Deflection of the stereocilia results in the opening of mechanoelectrical transduction ion channels residing on the stereocilia and ultimately triggers the initiation of a nerve impulse, which is relayed by the auditory nerve to the brain for interpretation (5).Normal development, functioning, and maintenance of hair cell structures depend on tight regulation of the hair cell cytoskeleton as it controls the height and width of the stereociliary bundles. Each stereocilium inserts into the second actin-rich structure, the cuticular plate, a web-like arrangement of actin filaments and associated actin-binding proteins located just beneath the basal surface of the stereocilia (3, 6). This fibrous matrix produces a foundation into which the stereocilia are anchored, and it furnishes physical support for maintaining the stereociliary bundles in their upright positions. Finally, the zonula adherens contains a thick band of actin filaments which encircles each hair cell. Its function is to both unite the hair cells with the surrounding matrix of epithelial cells and provide tension across the cuticular plate in a manner analogous to the springs found on a trampoline. Based on actin involvement in cochlear cell function, it is not surprising that mutations in the components of these filamentous networks cause deafness.In the majority of non-muscle cells in the body, the predominant actin isoform is β-nonmuscle actin with smaller amounts of γ-nonmuscle isoactin also present (7). These two isoforms differ at only 4 of the 375 amino acid residues in the protein. The N terminus of β-actin has three Asp residues, whereas that of γ-actin has three Glu residues. The other difference occurs at position 10; Val in β-actin is substituted by Ile in γ-actin. In the hair cell, however, γ-nonmuscle actin is the major form comprising about 75% of the total actin in the cell based on immunological studies with actin isoform-specific antibodies (8, 9).Six point mutations in the coding region of the γ-nonmuscle actin gene (ACT1G) have been shown to cause a subtype of autosomal dominant nonsyndromic progressive sensorineural hearing loss designated DNFA20/26 (1012). Patients carrying these mutations hear fine until their teens and twenties. Typically at this point the patients demonstrate marked decreases in their ability to hear high frequency sounds ranging from 4 to 8 kHz. As the patient ages the initial high frequency of hearing loss spreads over a broader range of frequencies (250 Hz to 8 kHz) until the overall loss becomes profound.Until recently nothing was known about how the biochemical effects of these actin mutations might ultimately lead to hearing loss. To address this problem, we devised a model system in which we introduced these six point mutations into their corresponding positions in yeast actin whose sequence is 90% identical to that of γ-nonmuscle isoactin. Furthermore, each of the mutation sites in γ-nonmuscle actin is identical to the corresponding site in yeast actin. This system allows for the analysis of these mutation effects in vivo and for the purification of enough of the mutant proteins to assess their biochemical and biophysical properties. Although these mutations produced allele-specific effects in yeast cells harboring them, the purified mutant actins, with one exception, V370A, displayed polymerization characteristics virtually identical to those of WT yeast actin (13). The lack of apparent gross polymerization defects associated with these mutant actins suggested that, instead, the deafness associated with them might result from misregulation of cytoskeletal dynamics because of altered interactions with one or more of the battery of actin-binding proteins known to control cytoskeletal function and dynamics.The actin filament is a polar structure (14, 15) with a barbed end, the preferred end for actin monomer addition, and a pointed end, which is the predominant end for monomer dissociation from the filament (16). This polarity is translated to each monomer, with the barbed end defined by subdomains 1 and 3 of the actin and the pointed end by subdomains 2 and 4 (Fig. 1). Four of the six mutations, K118M, T278I, P332A, and V370A, are in positions that could affect the barbed end surface, which interfaces with the pointed end of the next lower monomer in the filament strand (15). Moreover, three of the mutations, T89I and K118M and V370A, are predicted to be part of the actin surface comprising the tight and weak binding sites, respectively, for the actin-binding protein cofilin (17).Open in a separate windowFIGURE 1.Locations of the six γ-actin deafness mutations in yeast. A, front view of the crystal structure of the yeast actin monomer (63), modified from Protein Data Bank code 1YAG using Swiss-PdbViewer Version 4.01. The positions of the mutations are modeled in space-fill and color-highlighted as follows: T89I, orange; K118M, red; P264L, blue; T278I, green; P332A purple; cyan, V370A. ATP is modeled in ball-stick and colored in Corey-Pauling-Koltun. Numbers denote the actin subdomains, and N and C mark the respective termini. B, model illustrating the longitudinal contacts between two neighboring monomers within the same filament strand based on the Holmes filament model (15). Each monomer is labeled as in panel A with regard to the mutations, ADP, and subdomains. The red dashed circle denotes the general vicinity on the actin surface to which cofilin binds (21, 41, 42).Cofilin is a small, 15–21-kDa actin-binding protein that, depending on its concentration, can sequester actin monomers (18, 19), sever/disassemble actin filaments (2024), and promote filament nucleation by stabilizing spontaneously forming actin nuclei (25, 26). Cofilin plays an essential role in the regulation of actin cytoskeletal dynamics (18, 2731) and is found in cochlear cells (neibank.nei.nih.gov). Cofilin-dependent filament scission/disassembly results from its disruption of both lateral (24, 32) and longitudinal (21) contacts between neighboring monomers within the filament. Thus, an actin mutation which results in destabilization of monomer-monomer interfaces or alters the cofilin binding site might make the filament more susceptible to cofilin severing. An alternative function of cofilin that might be affected by these mutations is its ability to bind to either ATP monomers being added to the barbed end of the filament or ADP-actin monomers being released from the pointed end of the filament (23, 33, 34). Altered binding could lead to altered degrees of monomer sequestration by the cofilin ultimately affecting filament length and polymer dynamics.In this study we have assessed the ability of yeast actins carrying the deafness-causing actin mutations to interact with yeast cofilin in vitro. Based on our results in vitro, which demonstrated allele-specific altered actin-cofilin interactions, we then dampened cofilin function in vivo to determine whether altered regulation of the mutant actins by cofilin was responsible for the glycerol phenotypes associated with these mutations in vivo in yeast (13).  相似文献   

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

6.
7.
Dynamin2 GTPase and Cortactin Remodel Actin Filaments   总被引:1,自引:0,他引:1  
The large GTPase dynamin, best known for its activities that remodel membranes during endocytosis, also regulates F-actin-rich structures, including podosomes, phagocytic cups, actin comet tails, subcortical ruffles, and stress fibers. The mechanisms by which dynamin regulates actin filaments are not known, but an emerging view is that dynamin influences F-actin via its interactions with proteins that interact directly or indirectly with actin filaments. We show here that dynamin2 GTPase activity remodels actin filaments in vitro via a mechanism that depends on the binding partner and F-actin-binding protein, cortactin. Tightly associated actin filaments cross-linked by dynamin2 and cortactin became loosely associated after GTP addition when viewed by transmission electron microscopy. Actin filaments were dynamically unraveled and fragmented after GTP addition when viewed in real time using total internal reflection fluorescence microscopy. Cortactin stimulated the intrinsic GTPase activity of dynamin2 and maintained a stable link between actin filaments and dynamin2, even in the presence of GTP. Filaments remodeled by dynamin2 GTPase in vitro exhibit enhanced sensitivity to severing by the actin depolymerizing factor, cofilin, suggesting that GTPase-dependent remodeling influences the interactions of actin regulatory proteins and F-actin. The global organization of the actomyosin cytoskeleton was perturbed in U2-OS cells depleted of dynamin2, implicating dynamin2 in remodeling actin filaments that comprise supramolecular F-actin arrays in vivo. We conclude that dynamin2 GTPase remodels actin filaments and plays a role in orchestrating the global actomyosin cytoskeleton.Controlled assembly and disassembly of actin filaments underlies movement, shape, division, trafficking of lipids and proteins of the cell and pathogenesis by infectious bacteria and viruses. Several proteins and signaling circuits modulate actin filament dynamics, including proteins that nucleate formation of new filaments, filament cross-linking proteins that stabilize branched and bundled filament arrays, and depolymerizing factors that promote filament disassembly (1). Studies with reconstituted systems show that a single actin nucleating factor, such as the Arp2/3 complex together with a nucleation-promoting factor, a barbed end capping protein to preserve the actin monomer pool and promote nucleation, and a filament disassembly factor, such as ADF/cofilin, are sufficient to establish a dynamic dendritic actin network in vitro that mimics many properties of actin networks at the leading edge of migrating cells (24). However, the mechanisms for coordinating the organization and dynamics of actin filaments associated with higher-order cellular structures such as the subcortical F-actin network, F-actin at focal adhesions, and actomyosin arrays are not as well understood.Considerable evidence indicates that the large GTPase dynamin, a key mediator of membrane remodeling and fission, also influences actin filaments (reviewed in Refs. 57). Although the mechanisms are unknown, dynamin could influence actin filaments via its interactions with a number of proteins that directly or indirectly regulate actin filament assembly, filament stability, or filament organization. For example, several protein scaffolds biochemically link dynamin and the Arp2/3 complex activating factor, N-WASP, suggesting that the machinery for de novo actin assembly may be targeted or activated by dynamin (6, 8, 9). Dynamin2 is associated with several dynamic F-actin-containing structures in vivo, including podosomes, F-actin comet tails, phagocytic cups, dynamic cortical ruffles, and pedestal structures elaborated by enteropathogenic Escherichia coli (1020). Cortactin, which directly binds both dynamin and actin filaments, is associated with many of the same dynamic actin structures as dynamin (5, 7) and is required for both clathrin-dependent and -independent endocytosis (21, 22). Thus, dynamin-cortactin interaction may be an important link between actin filaments and dynamin during formation or turnover of F-actin-rich structures.Considerable evidence supports the notion that GTP hydrolysis by dynamin catalyzes membrane fission activity via GTPase-dependent changes in conformation (23, 24) or via GTPase-dependent cycles of assembly and disassembly (25, 26). We hypothesize that GTPase-dependent changes in dynamin linked via its interacting proteins to actin filaments or actin regulators could similarly influence actin filaments. Overexpressed, dominant negative dynamin mutant proteins impaired in binding or hydrolyzing GTP (most often the dynamin-K44A mutation) perturb a variety of F-actin-rich cellular structures, including stress fibers and focal adhesions (27, 28), dendritic spines of neurons (29), podosomes (12, 30), actin comet tails (13, 14), phagocytic cups and bacteria-induced pedestal structures (16, 19), and dynamic cortical ruffles (15, 17). In addition, F-actin of stress fibers and overall cell morphology were perturbed in Clone9 cells expressing a mutant dynamin2 protein lacking the C-terminal proline-rich domain, the domain through which dynamin2 interacts with actin regulatory factors (11). Whereas existing data indicates that the specific effects of dynamin GTPase activity on F-actin structures are cell type- and structure-specific, a general conclusion is that dynamin GTPase activity influences the organization or turnover of a subset of actin filaments.To determine the mechanisms by which dynamin2 GTPase activity influences actin filaments, we developed biochemical and microscopic approaches to quantitatively assess and observe GTPase-dependent effects on actin filaments formed in vitro with Arp2/3 complex, cortactin, and dynamin2. The activities of dynamin2 on actin filaments in vivo were examined in cells with disrupted dynamin2 function using siRNA2-mediated suppression or pharmacologic inhibition. We report that dynamin2 GTPase, together with cortactin, functions as a dynamic actin filament remodeling complex that influences the global organization of the actomyosin cytoskeleton.  相似文献   

8.
A decoding algorithm is tested that mechanistically models the progressive alignments that arise as the mRNA moves past the rRNA tail during translation elongation. Each of these alignments provides an opportunity for hybridization between the single-stranded, -terminal nucleotides of the 16S rRNA and the spatially accessible window of mRNA sequence, from which a free energy value can be calculated. Using this algorithm we show that a periodic, energetic pattern of frequency 1/3 is revealed. This periodic signal exists in the majority of coding regions of eubacterial genes, but not in the non-coding regions encoding the 16S and 23S rRNAs. Signal analysis reveals that the population of coding regions of each bacterial species has a mean phase that is correlated in a statistically significant way with species () content. These results suggest that the periodic signal could function as a synchronization signal for the maintenance of reading frame and that codon usage provides a mechanism for manipulation of signal phase.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]  相似文献   

9.
10.
Formins assemble non-branched actin filaments and modulate microtubule dynamics during cell migration and cell division. At the end of mitosis formins contribute to the generation of actin filaments that form the contractile ring. Rho small GTP-binding proteins activate mammalian diaphanous-related (mDia) formins by directly binding and disrupting an intramolecular autoinhibitory mechanism. Although the Rho-regulated activation mechanism is well characterized, little is known about how formins are switched off. Here we reveal a novel mechanism of formin regulation during cytokinesis based on the following observations; 1) mDia2 is degraded at the end of mitosis, 2) mDia2 is targeted for disposal by post-translational ubiquitin modification, 3) forced expression of activated mDia2 yields binucleate cells due to failed cytokinesis, and 4) the cytokinesis block is dependent upon mDia2-mediated actin assembly as versions of mDia2 incapable of nucleating actin but that still stabilize microtubules have no effect on cytokinesis. We propose that the tight control of mDia2 expression and ubiquitin-mediated degradation is essential for the completion of cell division. Because of the many roles for formins in cell morphology, we discuss the relevance of mDia protein turnover in other processes where ubiquitin-mediated proteolysis is an essential component.Formin proteins play a role in diverse processes such as cell migration (1, 2), vesicle trafficking (3, 4), tumor suppression (5, 6), and microtubule stabilization (7, 8). Formins also play an essential and conserved role in cytokinesis (911). Proper cell division is essential in all animals to maintain the integrity of their genome. Failure to complete cytokinesis can result in genomic instability and ultimately lead to disease such as cancer (12).The members of the mDia2 family of formins are autoregulated Rho effectors that remodel the cytoskeleton by nucleating and elongating non-branched actin filaments (13). The amino terminus of mDia contains a GTPase binding domain (GBD) that directs interaction with specific Rho small GTP-binding proteins. The adjacent Dia inhibitory domain (DID) mediates mDia autoregulation through its interaction with the carboxyl-terminal diaphanous autoregulatory domain (DAD) (14, 15). Between the DID and DAD domains lie the conserved formin homology 1 (FH1) and FH2 domains. The FH1 domain is a proline-rich region that mediates binding to other proteins such as profilin, Src, and Dia-interacting protein (1619). In contrast, the FH2 domain binds monomeric actin to generate filamentous actin (F-actin) and can also bind microtubules directly to induce their stabilization (8, 20).Although the mechanism of mDia activation is well characterized, little is known about its inactivation. Previous reports have suggested that formins can cycle between active, partially active, and inactive states (21, 22) due to GTP hydrolysis upon Rho binding to GTPase-activating proteins. Another formin inactivation mechanism is through mDia interactions with Dia-interacting protein (23). In the context of cortical actin assembly, Dia-interacting protein negatively regulates mDia2 actin polymerization but has no effect on mDia1 actin polymerization despite its ability to interact with both proteins directly (17). Because of the fundamental role for formins in cell division, we sought to identify how mDia2 is inactivated in mitosis.During cell division, the expression level and activity of many proteins (e.g. cyclins and Aurora and Polo kinases) are tightly regulated (24). A unifying regulatory mechanism among these proteins is ubiquitin-mediated proteolysis. In this study we find that mDia2 protein levels are constant from S phase into mitosis and dramatically decrease at the end of mitosis due to ubiquitin-mediated degradation. Failure to inhibit mDia2 actin assembly results in multinucleation, which supports an essential role for the tight regulation of mDia2 during cell division.  相似文献   

11.
Mitochondria play a central role in energy metabolism and cellular survival, and consequently mitochondrial dysfunction is associated with a number of human pathologies. Reversible protein phosphorylation emerges as a central mechanism in the regulation of several mitochondrial processes. In skeletal muscle, mitochondrial dysfunction is linked to insulin resistance in humans with obesity and type 2 diabetes. We performed a phosphoproteomics study of functional mitochondria isolated from human muscle biopsies with the aim to obtain a comprehensive overview of mitochondrial phosphoproteins. Combining an efficient mitochondrial isolation protocol with several different phosphopeptide enrichment techniques and LC-MS/MS, we identified 155 distinct phosphorylation sites in 77 mitochondrial phosphoproteins, including 116 phosphoserine, 23 phosphothreonine, and 16 phosphotyrosine residues. The relatively high number of phosphotyrosine residues suggests an important role for tyrosine phosphorylation in mitochondrial signaling. Many of the mitochondrial phosphoproteins are involved in oxidative phosphorylation, tricarboxylic acid cycle, and lipid metabolism, i.e. processes proposed to be involved in insulin resistance. We also assigned phosphorylation sites in mitochondrial proteins involved in amino acid degradation, importers and transporters, calcium homeostasis, and apoptosis. Bioinformatics analysis of kinase motifs revealed that many of these mitochondrial phosphoproteins are substrates for protein kinase A, protein kinase C, casein kinase II, and DNA-dependent protein kinase. Our results demonstrate the feasibility of performing phosphoproteome analysis of organelles isolated from human tissue and provide novel targets for functional studies of reversible phosphorylation in mitochondria. Future comparative phosphoproteome analysis of mitochondria from healthy and diseased individuals will provide insights into the role of abnormal phosphorylation in pathologies, such as type 2 diabetes.Mitochondria are the primary energy-generating systems in eukaryotes. They play a crucial role in oxidative metabolism, including carbohydrate metabolism, fatty acid oxidation, and urea cycle, as well as in calcium signaling and apoptosis (1, 2). Mitochondrial dysfunction is centrally involved in a number of human pathologies, such as type 2 diabetes, Parkinson disease, and cancer (3). The most prevalent form of cellular protein post-translational modifications (PTMs),1 reversible phosphorylation (46), is emerging as a central mechanism in the regulation of mitochondrial functions (7, 8). The steadily increasing numbers of reported mitochondrial kinases, phosphatases, and phosphoproteins imply an important role of protein phosphorylation in different mitochondrial processes (911).Mass spectrometry (MS)-based proteome analysis is a powerful tool for global profiling of proteins and their PTMs, including protein phosphorylation (12, 13). A variety of proteomics techniques have been developed for specific enrichment of phosphorylated proteins and peptides and for phosphopeptide-specific data acquisition techniques at the MS level (14). Enrichment methods based on affinity chromatography, such as titanium dioxide (TiO2) (1517), zwitterionic hydrophilic interaction chromatography (ZIC-HILIC) (18), immobilized metal affinity chromatography (IMAC) (19, 20), and ion exchange chromatography (strong anion exchange and strong cation exchange) (21, 22), have shown high efficiencies for enrichment of phosphopeptides (14). Recently, we demonstrated that calcium phosphate precipitation (CPP) is highly effective for enriching phosphopeptides (23). It is now generally accepted that no single method is comprehensive, but combinations of different enrichment methods produce distinct overlapping phosphopeptide data sets to enhance the overall results in phosphoproteome analysis (24, 25). Phosphopeptide sequencing by mass spectrometry has seen tremendous advances during the last decade (26). For example, MS/MS product ion scanning, multistage activation, and precursor ion scanning are effective methods for identifying serine (Ser), threonine (Thr), and tyrosine (Tyr) phosphorylated peptides (14, 26).A “complete” mammalian mitochondrial proteome was reported by Mootha and co-workers (27) and included 1098 proteins. The mitochondrial phosphoproteome has been characterized in a series of studies, including yeast, mouse and rat liver, porcine heart, and plants (19, 2831). To date, the largest data set by Deng et al. (30) identified 228 different phosphoproteins and 447 phosphorylation sites in rat liver mitochondria. However, the in vivo phosphoproteome of human mitochondria has not been determined. A comprehensive mitochondrial phosphoproteome is warranted for further elucidation of the largely unknown mechanisms by which protein phosphorylation modulates diverse mitochondrial functions.The percutaneous muscle biopsy technique is an important tool in the diagnosis and management of human muscle disorders and has been widely used to investigate metabolism and various cellular and molecular processes in normal and abnormal human muscle, in particular the molecular mechanism underlying insulin resistance in obesity and type 2 diabetes (32). Skeletal muscle is rich in mitochondria and hence a good source for a comprehensive proteomics and functional analysis of mitochondria (32, 33).The major aim of the present study was to obtain a comprehensive overview of site-specific phosphorylation of mitochondrial proteins in functionally intact mitochondria isolated from human skeletal muscle. Combining an efficient protocol for isolation of skeletal muscle mitochondria with several different state-of-the-art phosphopeptide enrichment methods and high performance LC-MS/MS, we identified 155 distinct phosphorylation sites in 77 mitochondrial phosphoproteins, many of which have not been reported before. We characterized this mitochondrial phosphoproteome by using bioinformatics tools to classify functional groups and functions, including kinase substrate motifs.  相似文献   

12.
13.
Loss or inactivation of BLM, a helicase of the RecQ family, causes Bloom syndrome, a genetic disorder with a strong predisposition to cancer. Although the precise function of BLM remains unknown, genetic data has implicated BLM in the process of genetic recombination and DNA repair. Previously, we demonstrated that BLM can disrupt the RAD51-single-stranded DNA filament that promotes the initial steps of homologous recombination. However, this disruption occurs only if RAD51 is present in an inactive ADP-bound form. Here, we investigate interactions of BLM with the active ATP-bound form of the RAD51-single-stranded DNA filament. Surprisingly, we found that BLM stimulates DNA strand exchange activity of RAD51. In contrast to the helicase activity of BLM, this stimulation does not require ATP hydrolysis. These data suggest a novel BLM function that is stimulation of the RAD51 DNA pairing. Our results demonstrate the important role of the RAD51 nucleoprotein filament conformation in stimulation of DNA pairing by BLM.Mutations of BLM helicase cause Bloom syndrome (BS),2 a rare autosomal disorder, which is associated with stunted growth, facial sun sensitivity, immunodeficiency, fertility defects, and a greatly elevated incidence of many types of cancer occurring at an early age (1). BLM belongs to the highly conserved family of RecQ helicases that are required for the maintenance of genome integrity in all organisms (2, 3). There are five RecQ helicases in humans; mutations in three of them, WRN, RECQ4, and BLM, have been associated with the genetic abnormalities known as Werner, Rothmund-Thomson, and Bloom syndrome, respectively (4, 5). The cells from BS patients display genomic instability; the hallmark of BS is an increase in the frequency of sister chromatid and interhomolog exchanges (1, 6). Because homologous recombination (HR) is responsible for chromosomal exchanges, it is thought that BLM helicase functions in regulating HR (79). Also, BLM helicase is required for faithful chromosome segregation (10) and repair of stalled replication forks (11, 12), the processes that are linked to HR (1315). BLM was found to interact physically with RAD51, a key protein of HR (16) that catalyzes the central steps in HR including the search for homology and the exchange of strands between homologous ssDNA and dsDNA sequences (17). In cells, BLM forms nuclear foci, a subset of which co-localize with RAD51. Interestingly, the extent of RAD51 and BLM co-localization increases in response to ionizing radiation, indicating a possible role of BLM in the repair of DNA double-strand breaks (16).Biochemical studies suggest that BLM may perform several different functions in HR. BLM was shown to promote the dissociation of HR intermediates (D-loops) (1820), branch migration of Holliday junctions (21), and dissolution of double Holliday junctions acting in a complex with TopoIIIα and BLAP75 (2224). BLM may also facilitate DNA synthesis during the repair process by unwinding the DNA template in front of the replication fork (25). In addition, BLM and its yeast homolog Sgs1 may play a role at the initial steps of DNA double-strand break repair by participating in exonucleolitic resection of the DNA ends to generate DNA molecules with the 3′-ssDNA tails, a substrate for RAD51 binding (2629).In vivo, the process of HR is tightly regulated by various mechanisms (30). Whereas some proteins promote HR (14, 31), others inhibit this process, thereby preventing its untimely initiation (32, 33). Disruption of the Rad51-ssDNA nucleoprotein filament appears to be an especially important mechanism of controlling HR. This filament disruption activity was demonstrated for the yeast Srs2 helicase (34, 35) and human RECQ5 helicase (36). Recently, we found that BLM can also catalyze disruption of the RAD51-ssDNA filament (25). This disruption only occurs if the filament is present in an inactive ADP-bound form, e.g. in the presence of Mg2+. Conversion of RAD51 into an active ATP-bound form, e.g. in the presence of Ca2+ (37), renders the filament resistant to BLM disruption (25). In this study, we analyze the interactions of BLM with an active ATP-bound RAD51-ssDNA filament. Surprisingly, we found that BLM stimulates the DNA strand exchange activity of RAD51. Thus, depending on the conformational state of the RAD51 nucleoprotein filament, BLM may either inhibit or stimulate the DNA strand exchange activity of RAD51. Our analysis demonstrated that, in contrast to several known stimulatory proteins that act by promoting formation of the RAD51-ssDNA filament, BLM stimulates the DNA strand exchange activity of RAD51 at a later stage, during synapsis. Stimulation appears to be independent of the ATPase activity of BLM. We suggest that this stimulation of RAD51 may represent a novel function of BLM in homologous recombination.  相似文献   

14.
15.
16.
A Boolean network is a model used to study the interactions between different genes in genetic regulatory networks. In this paper, we present several algorithms using gene ordering and feedback vertex sets to identify singleton attractors and small attractors in Boolean networks. We analyze the average case time complexities of some of the proposed algorithms. For instance, it is shown that the outdegree-based ordering algorithm for finding singleton attractors works in time for , which is much faster than the naive time algorithm, where is the number of genes and is the maximum indegree. We performed extensive computational experiments on these algorithms, which resulted in good agreement with theoretical results. In contrast, we give a simple and complete proof for showing that finding an attractor with the shortest period is NP-hard.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]  相似文献   

17.
The kinetochore, which consists of centromere DNA and structural proteins, is essential for proper chromosome segregation in eukaryotes. In budding yeast, Sgt1 and Hsp90 are required for the binding of Skp1 to Ctf13 (a component of the core kinetochore complex CBF3) and therefore for the assembly of CBF3. We have previously shown that Sgt1 dimerization is important for this kinetochore assembly mechanism. In this study, we report that protein kinase CK2 phosphorylates Ser361 on Sgt1, and this phosphorylation inhibits Sgt1 dimerization.The kinetochore is a structural protein complex located in the centromeric region of the chromosome coupled to spindle microtubules (1, 2). The kinetochore generates a signal to arrest cells during mitosis when it is not properly attached to microtubules, thereby preventing chromosome missegregation, which can lead to aneuploidy (3, 4). The molecular structure of the kinetochore complex of the budding yeast Saccharomyces cerevisiae has been well characterized; it is composed of more than 70 proteins, many of which are conserved in mammals (2).The centromere DNA in the budding yeast is a 125-bp region that contains three conserved regions, CDEI, CDEII, and CDEIII (5, 6). CDEIII (25 bp) is essential for centromere function (7) and is bound to a key component of the centromere, the CBF3 complex. The CBF3 complex contains four proteins, Ndc10, Cep3, Ctf13 (815), and Skp1 (14, 15), all essential for viability. Mutations in any of the CBF3 proteins abolish the ability of CDEIII to bind to CBF3 (16, 17). All of the kinetochore proteins, except the CDEI-binding Cbf1 (1820), localize to the kinetochores in a CBF3-dependent manner (2). Thus, CBF3 is a fundamental kinetochore complex, and its mechanism of assembly is of great interest.We have previously found that Sgt1 and Skp1 activate Ctf13; thus, they are required for assembly of the CBF3 complex (21). The molecular chaperone Hsp90 is also required to form the active Ctf13-Skp1 complex (22). Sgt1 has two highly conserved motifs that are required for protein-protein interaction: the tetratricopeptide repeat (21) and the CHORD protein and Sgt1-specific motif. We and others have found that both domains are important for the interaction of Sgt1 with Hsp90 (2326), which is required for assembly of the core kinetochore complex. This interaction is an initial step in kinetochore activation (24, 26, 27), which is conserved between yeast and humans (28, 29).We have recently shown that Sgt1 dimerization is important for Sgt1-Skp1 binding and therefore for kinetochore assembly (30). In this study, we have found that protein kinase CK2 phosphorylates Sgt1 at Ser361, and this phosphorylation inhibits Sgt1 dimerization. Therefore, CK2 appears to regulate kinetochore assembly negatively in budding yeast.  相似文献   

18.
CCN3 (NOV), a putative ligand for integrin receptors, is tightly associated with the extracellular matrix and mediates diverse cellular functions, including cell adhesion and proliferation. CCN3 has been shown to negatively regulate growth although it promotes migration in a cell type-specific manner. In this study, overexpression of CCN3 reduces growth and increases intercellular adhesion of breast cancer cells. Interestingly, CCN3 overexpression also led to the formation of multiple pseudopodia that are enriched in actin, CCN3, and vinculin. Breast cancer cells preincubated with exogenous CCN3 protein also induced the same phenotype, indicating that secreted CCN3 is sufficient to induce changes in cell morphology. Surprisingly, extracellular CCN3 is internalized to the early endosomes but not to the membrane protrusions, suggesting pseudopodia-enriched CCN3 may derive from a different source. The presence of an intracellular variant of CCN3 will be consistent with our finding that the cytoplasmic tail of the gap junction protein connexin43 (Cx43) associates with CCN3. Cx43 is a channel protein permitting intercellular communication to occur. However, neither the channel properties nor the protein levels of Cx43 are affected by the CCN3 protein. In contrast, CCN3 proteins are down-regulated in the absence of Cx43. Finally, we showed that overexpression of CCN3 increases the activity of the small GTPase Rac1, thereby revealing a pathway that links Cx43 directly to actin reorganization.The CCN (CYR61/Connective Tissue Growth Factor/Nephroblastoma Overexpressed) family of multimodular proteins mediates diverse cellular functions, including cell adhesion, migration, and proliferation (13). Overexpression of CCN3, one of the founding members of the family, inhibits proliferation in most types of tumors such as glioblastoma and Ewing sarcoma (4, 5). Similarly, down-regulation of CCN3 has been suggested to promote melanoma progression (6). On the other hand, CCN3 can also promote migration in sarcoma and glioblastoma (4, 7), although a separate study shows that it decreases the invasion of melanoma (6). Therefore, in contrast to its role in growth suppression, the role of CCN3 signaling in cell motility is less clear.Most evidence suggests CCN3 mediates its effects by binding to the integrin proteins, such as the αVβ3 receptors (8, 9), and that CCN3 alters cell adhesion in an integrin-dependent fashion (4, 10). In melanocytes, the discoidin domain receptor 1 mediates CCN3-dependent adhesion (11). CCN3 has also been observed to associate with Notch1 (12), fibulin 1C (13), S100A4 (14), and the gap junction protein Cx433 (15, 16), suggesting that CCN3 may also modulate cell growth via non-integrin signaling pathways.Gap junction proteins are best known for forming channels between cells, contributing to intercellular communication by allowing the exchange of small ions and molecules (17, 18). Consequently, attenuated intercellular communication has been implicated in promoting carcinogenesis (19, 20). Recent evidence has indicated that connexins can mediate channel-independent growth control through interaction of their C-terminal cytoplasmic tail with various intracellular signaling molecules (2123). In addition, many Cx43-interacting proteins, including ZO-1 (zonula occludens-1) (24), Drebrin (25), and N-cadherin (26) associate with F-actin, thus placing Cx43 in close proximity to the actin cytoskeleton.In this study, we show for the first time that CCN3 reorganizes the actin cytoskeleton of the breast cancer cells MDA-MB-231 with the formation of multiple cell protrusions, possibly by activating the small GTPase Rac1. Our results also suggest an alternative route by which Cx43 may be functionally linked to actin cytoskeletal signaling via CCN3.  相似文献   

19.
20.
SPA2 encodes a yeast protein that is one of the first proteins to localize to sites of polarized growth, such as the shmoo tip and the incipient bud. The dynamics and requirements for Spa2p localization in living cells are examined using Spa2p green fluorescent protein fusions. Spa2p localizes to one edge of unbudded cells and subsequently is observable in the bud tip. Finally, during cytokinesis Spa2p is present as a ring at the mother–daughter bud neck. The bud emergence mutants bem1 and bem2 and mutants defective in the septins do not affect Spa2p localization to the bud tip. Strikingly, a small domain of Spa2p comprised of 150 amino acids is necessary and sufficient for localization to sites of polarized growth. This localization domain and the amino terminus of Spa2p are essential for its function in mating. Searching the yeast genome database revealed a previously uncharacterized protein which we name, Sph1p (Spa2p homolog), with significant homology to the localization domain and amino terminus of Spa2p. This protein also localizes to sites of polarized growth in budding and mating cells. SPH1, which is similar to SPA2, is required for bipolar budding and plays a role in shmoo formation. Overexpression of either Spa2p or Sph1p can block the localization of either protein fused to green fluorescent protein, suggesting that both Spa2p and Sph1p bind to and are localized by the same component. The identification of a 150–amino acid domain necessary and sufficient for localization of Spa2p to sites of polarized growth and the existence of this domain in another yeast protein Sph1p suggest that the early localization of these proteins may be mediated by a receptor that recognizes this small domain.Polarized cell growth and division are essential cellular processes that play a crucial role in the development of eukaryotic organisms. Cell fate can be determined by cell asymmetry during cell division (Horvitz and Herskowitz, 1992; Cohen and Hyman, 1994; Rhyu and Knoblich, 1995). Consequently, the molecules involved in the generation and maintenance of cell asymmetry are important in the process of cell fate determination. Polarized growth can occur in response to external signals such as growth towards a nutrient (Rodriguez-Boulan and Nelson, 1989; Eaton and Simons, 1995) or hormone (Jackson and Hartwell, 1990a , b ; Segall, 1993; Keynes and Cook, 1995) and in response to internal signals as in Caenorhabditis elegans (Goldstein et al., 1993; Kimble, 1994; Priess, 1994) and Drosophila melanogaster (St Johnston and Nusslein-Volhard, 1992; Anderson, 1995) early development. Saccharomyces cerevisiae undergo polarized growth towards an external cue during mating and to an internal cue during budding. Polarization towards a mating partner (shmoo formation) and towards a new bud site requires a number of proteins (Chenevert, 1994; Chant, 1996; Drubin and Nelson, 1996). Many of these proteins are necessary for both processes and are localized to sites of polarized growth, identified by the insertion of new cell wall material (Tkacz and Lampen, 1972; Farkas et al., 1974; Lew and Reed, 1993) to the shmoo tip, bud tip, and mother–daughter bud neck. In yeast, proteins localized to growth sites include cytoskeletal proteins (Adams and Pringle, 1984; Kilmartin and Adams, 1984; Ford, S.K., and J.R. Pringle. 1986. Yeast. 2:S114; Drubin et al., 1988; Snyder, 1989; Snyder et al., 1991; Amatruda and Cooper, 1992; Lew and Reed, 1993; Waddle et al., 1996), neck filament components (septins) (Byers and Goetsch, 1976; Kim et al., 1991; Ford and Pringle, 1991; Haarer and Pringle, 1987; Longtine et al., 1996), motor proteins (Lillie and Brown, 1994), G-proteins (Ziman, 1993; Yamochi et al., 1994; Qadota et al., 1996), and two membrane proteins (Halme et al., 1996; Roemer et al., 1996; Qadota et al., 1996). Septins, actin, and actin-associated proteins localize early in the cell cycle, before a bud or shmoo tip is recognizable. How this group of proteins is localized to and maintained at sites of cell growth remains unclear.Spa2p is one of the first proteins involved in bud formation to localize to the incipient bud site before a bud is recognizable (Snyder, 1989; Snyder et al., 1991; Chant, 1996). Spa2p has been localized to where a new bud will form at approximately the same time as actin patches concentrate at this region (Snyder et al., 1991). An understanding of how Spa2p localizes to incipient bud sites will shed light on the very early stages of cell polarization. Later in the cell cycle, Spa2p is also found at the mother–daughter bud neck in cells undergoing cytokinesis. Spa2p, a nonessential protein, has been shown to be involved in bud site selection (Snyder, 1989; Zahner et al., 1996), shmoo formation (Gehrung and Snyder, 1990), and mating (Gehrung and Snyder, 1990; Chenevert et al., 1994; Yorihuzi and Ohsumi, 1994; Dorer et al., 1995). Genetic studies also suggest that Spa2p has a role in cytokinesis (Flescher et al., 1993), yet little is known about how this protein is localized to sites of polarized growth.We have used Spa2p green fluorescent protein (GFP)1 fusions to investigate the early localization of Spa2p to sites of polarized growth in living cells. Our results demonstrate that a small domain of ∼150 amino acids of this large 1,466-residue protein is sufficient for targeting to sites of polarized growth and is necessary for Spa2p function. Furthermore, we have identified and characterized a novel yeast protein, Sph1p, which has homology to both the Spa2p amino terminus and the Spa2p localization domain. Sph1p localizes to similar regions of polarized growth and sph1 mutants have similar phenotypes as spa2 mutants.  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