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.
2.
Inactivating mutations of the neurofibromatosis 2 (NF2) gene, NF2, result predominantly in benign neurological tumors, schwannomas and meningiomas, in humans; however, mutations in murine Nf2 lead to a broad spectrum of cancerous tumors. The tumor-suppressive function of the NF2 protein, merlin, a membrane-cytoskeleton linker, remains unclear. Here, we identify the mammalian target of rapamycin complex 1 (mTORC1) as a novel mediator of merlin''s tumor suppressor activity. Merlin-deficient human meningioma cells and merlin knockdown arachnoidal cells, the nonneoplastic cell counterparts of meningiomas, exhibit rapamycin-sensitive constitutive mTORC1 activation and increased growth. NF2 patient tumors and Nf2-deficient mouse embryonic fibroblasts demonstrate elevated mTORC1 signaling. Conversely, the exogenous expression of wild-type merlin isoforms, but not a patient-derived L64P mutant, suppresses mTORC1 signaling. Merlin does not regulate mTORC1 via the established mechanism of phosphoinositide 3-kinase-Akt or mitogen-activated protein kinase/extracellular signal-regulated kinase-mediated TSC2 inactivation and may instead regulate TSC/mTOR signaling in a novel fashion. In conclusion, the deregulation of mTORC1 activation underlies the aberrant growth and proliferation of NF2-associated tumors and may restrain the growth of these lesions through negative feedback mechanisms, suggesting that rapamycin in combination with phosphoinositide 3-kinase inhibitors may be therapeutic for NF2.Meningiomas are mesenchymal tumors that arise from the arachnoid layer covering the brain and spinal cord and account for approximately 30% of all primary intracranial neoplasms (30). Most sporadic meningiomas (60%) display somatic inactivation of the NF2 gene. Germ line mutations of NF2 are associated with neurofibromatosis 2 (NF2), a dominantly inherited disorder characterized by multiple nervous system tumors, including schwannomas and meningiomas (33). Although most meningiomas are benign (WHO grade I), they often cause significant morbidity due to compression of the adjacent brain or spinal cord. Benign meningiomas also have recurrence rates of up to 20% over 10 years. Ten percent of meningiomas are classified as atypical (WHO grade II) or anaplastic (WHO grade III) and display more aggressive clinical behavior, with rapid growth and increased recurrence rates (6, 21). The current standard of care is maximal surgical resection, with adjuvant radiation reserved for progressive tumors or those with aggressive features (e.g., WHO grade II or III). The treatment strategy for meningiomas that progress despite surgery and radiation remains limited, and currently there is no effective chemotherapy.The development of effective therapies has been hampered, in part, by our incomplete understanding of the signals influencing meningioma cell growth. Enhanced expression of certain peptide and steroid growth factors and receptors in meningioma tissue suggests that specific autocrine growth-stimulatory loops may be functionally important in meningioma cell proliferation (20, 38). The scarcity of established meningioma models that would allow for the assessment of growth-regulatory mechanisms has also hampered progress. Recently, we have developed reliable meningioma models that overcome the challenges of the low growth rates and senescence of primary benign meningioma cells (19).Biallelic inactivation of the NF2 gene is detected in the majority of sporadic meningiomas and nearly all schwannomas (11). The tumor suppressor gene NF2 encodes merlin (also called schwannomin), a member of the ezrin-radixin-moesin (ERM) protein family that functions to link membrane proteins to the cortical actin cytoskeleton (31, 41). Like the ERM proteins, merlin has been implicated in the regulation of membrane organization and cytoskeleton-based cellular processes such as adhesion, migration, cell-cell contact, spreading, proliferation, and signal transduction (27). The loss of contact-dependent inhibition of proliferation is seen in several types of NF2-deficient cells (23, 29). Merlin controls cell proliferation in response to cell contact via CD44 (28) and functions together with the related tumor suppressor Expanded via the Hippo/Mst pathway in both Drosophila and some types of mammalian cells (14, 49). Although merlin is implicated in a wide range of cellular activities, the precise mechanism by which merlin mediates growth-inhibitory functions in human arachnoidal and Schwann cells and the way in which its loss results in tumor formation in NF2 remain poorly understood.We recently reported that primary human merlin-deficient meningioma cells exhibit a striking, enlarged-cell phenotype compared to nonneoplastic arachnoidal cell counterparts derived from the same patient (19). Interestingly, the tuberous sclerosis complex (TSC) tumor suppressor syndrome is characterized by widespread benign tumors that possess abnormally large cells (22). Mutations in the tumor suppressor genes TSC1 and TSC2 result in TSC syndrome, and the corresponding protein products, hamartin and tuberin (referred to as TSC1 and TSC2), function together as a complex that potently inhibits mammalian target of rapamycin complex 1 (mTORC1) (17). mTOR is an evolutionarily conserved Ser/Thr kinase that exists in one of two distinct functional complexes, TORC1 and TORC2. TORC1, which regulates autophagy, protein translation, and ribosome biogenesis, is potently and specifically inhibited by rapamycin (10, 46). TORC2, which is less sensitive to rapamycin, is important for cytoskeletal regulation and Akt/protein kinase B activation (16, 18, 36).The TSC1-TSC2 complex inhibits mTORC1 by acting as a GTPase-activating protein for the small GTPase Rheb (Ras homolog enriched in brain). Inactivation of the TSC1-TSC2 complex results in the accumulation of GTP-bound Rheb, which activates mTORC1 (10). In addition to naturally occurring mutations in the TSC1 and TSC2 genes, growth factor stimulation of the phosphoinositide 3-kinase (PI3K)-Akt pathway, as well as Ras/mitogen-activated protein kinase (MAPK) pathways, leads to the phosphorylation and inactivation of the TSC1-TSC2 complex and consequent activation of mTORC1 (17). The activation of mTORC1 results in the phosphorylation of two well-characterized effectors, eukaryotic translation initiation factor 4E-binding protein 1 (4EBP1) and S6 kinase 1 (S6K1), leading to an increase in ribosomal biogenesis and the selective translation of specific mRNA populations. As a critical regulator of cell growth and proliferation, the mTORC1 pathway is dysregulated in several hamartoma syndromes, as well as in many cancers (10).In this report, we identify the NF2 tumor suppressor protein, merlin, as a novel negative regulator of the mTORC1 pathway to control cell growth (cell size). We show that mTORC1 is constitutively activated in merlin-deficient human meningioma cells, leading to increased cell size. Furthermore, we suggest that the slow growth of merlin-deficient meningioma cells is due to a rapamycin-sensitive, mTORC1-S6K-dependent negative feedback loop that diminishes PI3K-Akt signaling in response to growth factor stimulation. The findings of these studies provide insight into the mechanism of merlin tumor suppressor activity and, moreover, indicate that rapamycin or rapamycin analogs in combination with PI3K inhibitors may provide promise as new therapeutics in the treatment of meningiomas and schwannomas.  相似文献   

3.
Signaling via the Pyk2-Src-Cbl complex downstream of integrins contributes to the assembly, organization, and dynamics of podosomes, which are the transient adhesion complexes of highly motile cells such as osteoclasts and dendritic cells. We previously demonstrated that the GTPase dynamin is associated with podosomes, regulates actin flux in podosomes, and promotes bone resorption by osteoclasts. We report here that dynamin associates with Pyk2, independent of dynamin''s GTPase activity, and reduces Pyk2 Y402 phosphorylation in a GTPase-dependent manner, leading to decreased Src binding to Pyk2. Overexpressing dynamin decreased the macrophage colony-stimulating factor- and adhesion-induced phosphorylation of Pyk2 in osteoclastlike cells, suggesting that dynamin is likely to regulate Src-Pyk2 binding downstream of integrins and growth factor receptors with important cellular consequences. Furthermore, catalytically active Src promotes dynamin-Pyk2 association, and mutating specific Src-phosphorylated tyrosine residues in dynamin blunts the dynamin-induced decrease in Pyk2 phosphorylation. Thus, since Src binds to Pyk2 through its interaction with phospho-Y402, our results suggest that Src activates a negative-feedback loop downstream of integrin engagement and other stimuli by promoting both the binding of dynamin to Pyk2-containing complexes and the dynamin-dependent decrease in Pyk2 Y402 phosphorylation, ultimately leading to the dissociation of Src from Pyk2.Podosomes are specialized transient actin-containing adhesion structures (11, 14, 37, 60) that are found in highly motile cells, such as osteoclasts, macrophages, dendritic cells, transformed metastatic cells, and v-src-transformed cells (37, 43), where they are thought to play important roles in cellular migration and invasion (34). In resorbing osteoclasts on bone, podosomes are concentrated within the sealing zone, a beltlike actin-rich structure that is important for adhesion and which delineates the resorptive region of the cell known as the ruffled border. Unlike focal adhesions, which are relatively stable structures (11, 60), the assembly and disassembly of podosomes occurs within minutes (t1/2 = 2 to 4 min) and involves the recruitment and activation of integrins, signaling proteins and scaffolding proteins (11, 14, 35, 47, 60). However, the mechanisms of action of key signaling proteins involved in podosome assembly and disassembly are only partially understood.The focal adhesion kinase Pyk2 has been linked to the proliferation, migration, and activity of a variety of mesenchymal, epithelial, and hematopoietic cell types. Several groups, including our own, have reported the importance of Pyk2 in podosome belt organization, cell spreading, and bone-resorbing activity in osteoclasts (18, 26, 31, 40, 65, 66). Pyk2 is recruited to activated β2 and β3 integrins (9, 20) at adhesion sites and is autophosphorylated at Y402 (17, 47, 50) via an intermolecular trans-acting mechanism (46). Although Pyk2 is partially activated by integrin-induced Ca2+ signaling (20, 50), the induction of Pyk2''s full catalytic activity requires the binding of Src via its SH2 domain to autophosphorylated Pyk2 Y402 and the subsequent phosphorylation of Pyk2 at functionally distinct sites, including Y579, Y580, and Y881 (17, 31, 46). The binding of Src to phosphorylated Pyk2, which leads to the formation of a multiprotein signaling complex at adhesion sites (17, 40, 50), is critical for Pyk2 activity, as demonstrated by the fact that Pyk2 phosphorylation and activity are significantly reduced in osteoclasts derived from Src−/− mice (17, 40). Src−/− osteoclasts also exhibit decreased motility (50) and decreased bone-resorbing activity (40, 54, 59), and we recently demonstrated that Src promotes both podosome formation and disassembly, as well as actin flux into existing podosomes and the organization of podosomes into a peripheral belt in osteoclasts (15).We have also demonstrated that the GTP-hydrolyzing protein dynamin-2, which is ubiquitously expressed and well known for its role in endocytosis (53), regulates actin remodeling in the podosomes of osteoclasts and Rous sarcoma virus-transformed baby hamster kidney cells (43). In addition, a dynamin-2 mutant that binds GTP with reduced affinity (dynK44A) (12) decreased the flux of actin into podosomes (43) and disrupted podosome belt formation in osteoclasts, thereby affecting osteoclast migration and bone-resorbing activity (8). The dynamin proteins, of which there are three homologous isoforms (3), contain several protein domains: a GTP-hydrolyzing domain (GTPase), a plextrin homology domain that mediates binding to phosphoinositides, a GTPase effector domain (GED), and a C-terminal proline-rich domain (PRD) (38, 45, 55) through which dynamin binds a number of functionally diverse SH3-containing molecules, such as Src, cortactin, Grb2, and N-Wasp (1, 7, 27, 39, 58). We previously reported that dynamin-2 partially colocalizes and associates with the E3-ubiquitin ligase Cbl within the podosome belt/sealing zone of osteoclasts, as well as in SYF cells, which lack the Src family kinases Src, Yes, and Fyn, and in HEK 293 cells that stably express the vitronectin receptor (293VnR) (8). Protein complexes containing dynamin-2 and Cbl, which are both substrates of Src (1, 2, 23, 50, 56), were disrupted in the presence of activated Src and stabilized in the absence of Src (8), demonstrating a key role of Src in regulating the formation of signaling complexes in osteoclasts downstream of integrins.In the present study, we sought to determine whether dynamin, which regulates podosome actin dynamics and bone resorption in osteoclasts, also associates with Pyk2 and/or regulates Pyk2''s activities in osteoclasts. We report here that dynamin associates with Pyk2 and promotes the dephosphorylation of Pyk2 Y402 and that catalytically active Src promotes both dynamin''s association with Pyk2 and the dynamin-induced dephosphorylation of Pyk2 Y402, resulting, in turn, in the decreased binding of Src to Pyk2. Thus, we propose that dynamin regulates podosome dynamics and osteoclast bone-resorbing activity by promoting the disassembly of the Pyk2-Src-Cbl complex that is formed in osteoclasts downstream of β3 integrin activation.  相似文献   

4.
Receptor protein tyrosine phosphatase α (RPTPα) is the mitotic activator of the protein tyrosine kinase Src. RPTPα serine hyperphosphorylation was proposed to mediate mitotic activation of Src. We raised phosphospecific antibodies to the two main serine phosphorylation sites, and we discovered that RPTPα Ser204 was almost completely dephosphorylated in mitotic NIH 3T3 and HeLa cells, whereas Ser180 and Tyr789 phosphorylation were only marginally reduced in mitosis. Concomitantly, Src pTyr527 and pTyr416 were dephosphorylated, resulting in 2.3-fold activation of Src in mitosis. Using inhibitors and knockdown experiments, we demonstrated that dephosphorylation of RPTPα pSer204 in mitosis was mediated by PP2A. Mutation of Ser204 to Ala did not activate RPTPα, and intrinsic catalytic activity of RPTPα was not affected in mitosis. Interestingly, binding of endogenous Src to RPTPα was induced in mitosis. GRB2 binding to RPTPα, which was proposed to compete with Src binding to RPTPα, was only modestly reduced in mitosis, which could not account for enhanced Src binding. Moreover, we demonstrate that Src bound to mutant RPTPα-Y789F, lacking the GRB2 binding site, and mutant Src with an impaired Src homology 2 (SH2) domain bound to RPTPα, illustrating that Src binding to RPTPα is not mediated by a pTyr-SH2 interaction. Mutation of RPTPα Ser204 to Asp, mimicking phosphorylation, reduced coimmunoprecipitation with Src, suggesting that phosphorylation of Ser204 prohibits binding to Src. Based on our results, we propose a new model for mitotic activation of Src in which PP2A-mediated dephosphorylation of RPTPα pSer204 facilitates Src binding, leading to RPTPα-mediated dephosphorylation of Src pTyr527 and pTyr416 and hence modest activation of Src.Protein tyrosine phosphatases (PTPs) are responsible for dephosphorylation of the phosphotyrosyl residues. The human genome contains approximately 100 genes that encode members of the four PTP families, and most of them have mouse orthologues (2, 48). According to their subcellular localization, the classical PTPs, encoded by less than half of the total PTP genes, are divided into two subfamilies: cytoplasmic and receptor protein tyrosine phosphatases (RPTPs). The majority of the RPTPs contain, besides a variable extracellular domain and a transmembrane domain, two highly homologous phosphatase domains (27), with the membrane-proximal domain comprising most of the catalytic activity (33).RPTPα is a typical RPTP with a small, highly glycosylated extracellular domain (13). RPTPα function is regulated by many mechanisms, including proteolysis (18), oxidation (55), dimerization (7, 23, 24, 47, 52), and phosphorylation of serine and tyrosine residues (16, 17, 49). RPTPα is broadly expressed in many cell types, and over the years, RPTPα has been shown to be involved in a number of signaling mechanisms, including neuronal (15) and skeletal muscle (34) cell differentiation, neurite elongation (8, 9, 56), insulin receptor signaling downregulation (3, 28, 30, 31, 35), insulin secretion (25), activation of voltage-gated potassium channel Kv1.2 (51), long-term potentiation in hippocampal neurons (32, 38), matrix-dependent force transduction (53), and cell spreading and migration (21, 45, 57).The majority of the roles played in these cellular processes involve RPTPα''s ability to activate the proto-oncogenes Src and Fyn by dephosphorylating their C-terminal inhibitory phosphotyrosine (5, 15, 39, 45, 61). Normally, this phosphotyrosine (pTyr527 in chicken Src) binds to the Src homology 2 (SH2) domain, keeping the protein in an inactive closed conformation. A displacement mechanism was proposed for RPTPα-mediated Src activation in which pTyr789 of RPTPα is required to bind the SH2 domain of Src before RPTPα dephosphorylates Tyr527 (58). This model is the subject of debate since other studies show that RPTPα lacking Tyr789 is still able to dephosphorylate and activate Src (12, 26, 29, 56). In normal cells, Src reaches its activation peak during mitosis (4, 11, 40, 42), and with the help of overexpressing cells, it was shown that this activation is triggered mainly by RPTPα. The model that emerged is that RPTPα is activated in mitosis due to serine hyperphosphorylation and detaches from the GRB2 scaffolding protein (59, 60) that normally binds most of the pTyr789 of RPTPα via its SH2 domain (14, 17, 46). Two serine phosphorylation sites were mapped in the juxtamembrane domain of RPTPα, Ser180 and Ser204 (49). The kinases that were found responsible for their phosphorylation were protein kinase C delta (PKCdelta) (10) and CaMKIIalpha (9), but there is no clear evidence that these kinases are activated in mitosis. We set out to investigate the role of serine phosphorylation of RPTPα in mitotic activation of Src.We generated phosphospecific antibodies and show that RPTPα pSer204, but not pSer180, is dephosphorylated in mitotic NIH 3T3 and HeLa cells, concomitantly with activation of Src. Selective inhibitors suggested that PP2A was the phosphatase that dephosphorylated pSer204. RNA interference (RNAi)-mediated knockdown of the catalytic subunit of PP2A demonstrated that indeed PP2A was responsible for mitotic dephosphorylation of RPTPα pSer204. It is noteworthy that PP2A is known to be activated in mitosis. Intrinsic PTP activities of RPTPα were similar in unsynchronized and mitotic cells, and mutation of Ser204 did not activate RPTPα in in vitro PTP assays. Yet, Src binding to RPTPα was induced in mitotic NIH 3T3 cells and RPTPα-S204D with a phosphomimicking mutation at Ser204 coimmunoprecipitated less efficiently with Src. Based on our results, we propose a mechanism for mitotic activation of Src that is triggered by dephosphorylation of RPTPα pSer204, resulting in enhanced affinity for Src and subsequent dephosphorylation and activation of Src.  相似文献   

5.
Neurofibromatosis type 2 is an inherited autosomal disorder caused by biallelic inactivation of the NF2 tumor suppressor gene. The NF2 gene encodes a 70-kDa protein, merlin, which is a member of the ezrin-radixin-moesin (ERM) family. ERM proteins are believed to be regulated by a transition between a closed conformation, formed by binding of their N-terminal FERM domain and C-terminal tail domain (CTD), and an open conformation, in which the two domains do not interact. Previous work suggests that the tumor suppressor function of merlin is similarly regulated and that only the closed form is active. Therefore, understanding the mechanisms that control its conformation is crucial. We have developed a series of probes that measures merlin conformation by fluorescence resonance energy transfer, both as purified protein and in live cells. Using these tools, we find that merlin exists predominately as a monomer in a stable, closed conformation that is mediated by the central α-helical domain. The contribution from the FERM-CTD interaction to the closed conformation appears to be less important. Upon phosphorylation or interaction with an effector protein, merlin undergoes a subtle conformational change, suggesting a novel mechanism that modulates the interaction between the FERM domain and the CTD.Neurofibromatosis type 2 is an inherited autosomal disorder that is characterized by bilateral schwannomas of the eighth cranial nerve. The tumor suppressor gene responsible for this disorder, NF2, was cloned in 1993 (45). Biallelic inactivation of the NF2 gene is also seen in spontaneous schwannoma, meningioma, and malignant mesothelioma (22). In mouse models, deletion of the Nf2 gene is embryonic lethal, indicating an essential role for NF2 in development (24). Heterozygous mice develop a variety of aggressive metastatic tumors that have lost the wild-type allele (23). Targeted deletion of the Nf2 gene in Schwann cells leads to schwannoma formation (7). In vitro, Nf2-null cells grow to significantly higher densities (31), suggesting that contact inhibition of growth is impaired in these cells and that mediation of growth arrest at high cell density may be the basis for the tumor suppressor function of the NF2 gene. In normal fibroblasts, merlin is inactive as a growth suppressor in subconfluent cells, becoming activated as they approach confluence, thereby effecting contact inhibition of growth (26).The NF2 gene encodes a 70-kDa protein called merlin (for moesin, ezrin, radixin-like protein), which shares significant homology with members of the ezrin-radixin-moesin (ERM) branch of the Band 4.1 superfamily (45). The domain structure of merlin, also shared with other ERM proteins, consists of an N-terminal FERM domain, followed by a central α-helical region (CH) and a C-terminal tail domain (CTD). The merlin FERM domain has relatively high sequence similarity with other ERM family members, a 60 to 70% identity over the first 300 amino acids. The CH domain and the CTD show much lower identity (28 to 36%); however, the α-helical character of the CH domain is preserved, as is the heptad repeat pattern typical of α-helices that form coiled coils (46).The critical point of regulation of all the ERM proteins is a high-affinity intramolecular interaction between the C-terminal domain and the FERM domain (4) (Fig. (Fig.1).1). The FERM domain folds into a three-lobed cloverleaf structure that acts as a multifaceted docking site for protein binding partners (16, 39). The CTD, consisting of four major and two minor helices and a beta sheet, binds to the FERM domain by extending across the face of the F2 and F3 lobes (32). This intramolecular head-to-tail binding results in a “closed” conformation, with the C-terminal domain covering much of the surface of the FERM domain (32, 44). For ezrin, radixin and moesin, the CTD functions as a mask, blocking access of effector molecules, such as the cell surface receptors CD44 and ICAM2 and adaptor molecules, like EBP50/NHERF, to sites on the surface of the FERM domain (11, 25, 44). The interaction between the CTD and FERM domain is regulated by phosphatidyl inositol-(4,5)-bisphosphate (PIP2) binding to the FERM domain and by phosphorylation of a critical residue in the CTD (3, 6, 10, 49). This residue, threonine 567 in ezrin, is conserved throughout the ERM family (21). Phosphorylation introduces a negative charge and a bulky side group that effectively reduces the affinity of the interaction, releasing the CTD from the FERM domain and causing a transition to an open conformation. Low-angle rotary shadowing electron microscopy (13) and biochemical studies (12) of purified radixin suggest that in the open conformation it is an extended filamentous structure with globular N and C termini that is greater than 240 Å in length. Signal transduction systems, such as the epidermal growth factor and Rho A pathways, induce phosphorylation of ERM proteins at the conserved C-terminal threonine via a number of kinases, including Rho kinase and protein kinase Cα (21, 28). Thus, conformational regulation of ERM proteins can be a point of integration of ERM activity with signal transduction pathways. The overall concept of ERM regulation, then, is centered upon a transition between an inactive, closed conformation that is mediated by the FERM-CTD interaction and an active, open conformation that is regulated by phosphorylation. In these two states, ERM proteins likely interact with different sets of binding partners, resulting in distinct functional outcomes.Open in a separate windowFIG. 1.ERM tertiary structure as represented by the crystal structure of full-length Sf-moesin (20), but with the merlin amino acid sequence substituted for Sf-moesin. Approximate boundary amino acid residues for all domains appear at the top of the figure. Each domain is assigned a different color. The ERM structure consists of an N-terminal FERM domain folded into three lobes, F1, F2, and F3. This is followed by a central α-helical domain containing three subhelices (αA, αB, and αC) and a CTD with four short helices. An ERM protein is thought to have an open conformation, an extended structure with the FERM domain and the CTD separated by the α-helical domain, that is more than 240 Å long. In the closed conformation, the α-helical domain bends at the αA-αB junction and again at the αB-αC junction, causing the CTD to be positioned over F2 and F3 of the FERM domain. More than half of the surface of the FERM domain is masked by interaction with the CTD, αA, and parts of αB and αC.Like the classical ERMs, merlin is also thought to be regulated by changes in conformation. The FERM domain and the CTD of merlin interact with each other, albeit at a lower level of affinity than the ezrin FERM domain and the CTD (29). There are important differences, however, between merlin and the other ERM proteins. First, phosphorylation of the conserved C-tail threonine T576 has not been reported to occur in mammalian merlin, and nonphosphorylatable and phosphomimetic substitutions at this site have no effect on merlin activity (42). Instead, merlin is phosphorylated at serine 518 in the CTD, a target of the p21-activated kinase PAK and protein kinase A (1, 18, 47). The growth-suppressive function of merlin is activated by dephosphorylation of S518 by the phosphatase PP1δ in a density-dependent manner (14). Second, it was reported in a study using FERM domain and CTD truncates of merlin that only cotransfection of both the N-and C-terminal halves resulted in growth suppression (38). Together, these observations suggested a model of inactive, phosphorylated merlin in an open conformation that, upon cell-to-cell contact, is dephosphorylated and transitions to a closed, growth suppressive conformation.The existing model for conformational regulation described above is inferred from indirect data and assays that generally measure the interaction of isolated FERM and CTD truncates rather than full-length molecules (9, 29, 38). It has been impossible to test directly because tools have not been available to specifically assay for either the open or the closed form of merlin. Therefore, we have developed a series of probes that measures merlin conformation by fluorescence resonance energy transfer (FRET), both as purified protein and in live cells. Using these tools, we show that merlin exists predominately as a monomer in a stable, largely closed conformation. Additionally, we find that the closed conformation is largely mediated by the central α-helical domain; the contribution of the FERM-CTD interaction appears to be less significant than previously thought. Finally, we find that phosphorylation and protein interaction cause unexpectedly small changes in merlin conformation. We propose a new and more refined model for merlin regulation, in which merlin function is regulated by specific but subtle conformational changes that modulate the interaction between the FERM domain and the CTD.  相似文献   

6.
7.
Protein tyrosine kinase 6 (PTK6) is a nonmyristoylated Src-related intracellular tyrosine kinase. Although not expressed in the normal mammary gland, PTK6 is expressed in a majority of human breast tumors examined, and it has been linked to ErbB receptor signaling and AKT activation. Here we demonstrate that AKT is a direct substrate of PTK6 and that AKT tyrosine residues 315 and 326 are phosphorylated by PTK6. Association of PTK6 with AKT occurs through the SH3 domain of PTK6 and is enhanced through SH2 domain-mediated interactions following tyrosine phosphorylation of AKT. Using Src, Yes, and Fyn null mouse embryonic fibroblasts (SYF cells), we show that PTK6 phosphorylates AKT in a Src family kinase-independent manner. Introduction of PTK6 into SYF cells sensitized these cells to physiological levels of epidermal growth factor (EGF) and increased AKT activation. Stable introduction of active PTK6 into SYF cells also resulted in increased proliferation. Knockdown of PTK6 in the BPH-1 human prostate epithelial cell line led to decreased AKT activation in response to EGF. Our data indicate that in addition to promoting growth factor receptor-mediated activation of AKT, PTK6 can directly activate AKT to promote oncogenic signaling.Protein tyrosine kinase 6 (PTK6; also known as the breast tumor kinase BRK) is an intracellular Src-related tyrosine kinase (9, 48). Human PTK6 was identified in cultured human melanocytes (32) and breast tumor cells (39), while its mouse orthologue was cloned from normal small intestinal epithelial cell RNA (50). Although PTK6 shares overall structural similarity with Src family tyrosine kinases, it lacks an N-terminal myristoylation consensus sequence for membrane targeting (39, 51). As a consequence, PTK6 is localized to different cellular compartments, including the nucleus (14, 15). PTK6 is expressed in normal differentiated epithelial cells of the gastrointestinal tract (34, 42, 51), prostate (14), and skin (51-53). Expression of PTK6 is upregulated in different types of cancers, including breast carcinomas (6, 39, 54), colon cancer (34), ovarian cancer (47), head and neck cancers (33), and metastatic melanoma cells (16). The significance of apparent opposing signaling roles for PTK6 in normal differentiation and cancer is still poorly understood.In human breast tumor cells, PTK6 enhances signaling from members of the ErbB receptor family (10, 29, 30, 36, 40, 49, 54). In the HB4a immortalized human mammary gland luminal epithelial cell line, PTK6 promoted epidermal growth factor (EGF)-induced ErbB3 tyrosine phosphorylation and AKT activation (29). In response to EGF stimulation, PTK6 promoted phosphorylation of the focal adhesion protein paxillin and Rac1-mediated cell migration (10). PTK6 can be activated by the ErbB3 ligand heregulin and promotes activation of extracellular signal-regulated kinase 5 (ERK5) and p38 mitogen-activated protein kinase (MAPK) in breast cancer cells (40). PTK6 can also phosphorylate p190RhoGAP-A and stimulate its activity, leading to RhoA inactivation and Ras activation and thereby promoting EGF-dependent breast cancer cell migration and proliferation (49). Expression of PTK6 has been correlated with ErbB2 expression in human breast cancers (4, 5, 54).AKT (also called protein kinase B) is a serine-threonine kinase that is activated downstream of growth factor receptors (38). It is a key player in signaling pathways that regulate energy metabolism, proliferation, and cell survival (7, 45). Aberrant activation of AKT through diverse mechanisms has been discovered in different cancers (2). AKT activation requires phosphorylation of AKT on threonine residue 308 and serine residue 473. The significance of phosphorylation of AKT on tyrosine residues is less well understood. Src has been shown to phosphorylate AKT on conserved tyrosine residues 315 and 326 near the activation loop (11). Substitution of these two tyrosine residues with phenylalanine abolished AKT kinase activity stimulated by EGF (11). Use of the Src family inhibitor PP2 impaired AKT activation following IGF-1 stimulation of oligodendrocytes (13). The RET/PTC receptor tyrosine kinase that responds to glial cell-line-derived neurotrophic factor also phosphorylated AKT tyrosine residue 315 promoting activation of AKT (28). AKT tyrosine residue 474 was phosphorylated when cells were treated with the tyrosine phosphatase inhibitor pervanadate, and phosphorylation of tyrosine 474 contributed to full activation of AKT (12). Recently, the nonreceptor tyrosine kinase Ack1 was shown to regulate AKT tyrosine phosphorylation and activation (37).Here we show that AKT is a cytoplasmic substrate of the intracellular tyrosine kinase PTK6. We identify the tyrosine residues on AKT that are targeted by PTK6, and we demonstrate that tyrosine phosphorylation plays a role in regulating association between PTK6 and AKT. In addition, we show that PTK6 promotes AKT activation and cell proliferation in a Src-independent manner.  相似文献   

8.
9.
10.
11.
12.
13.
14.
Saccharomyces cerevisiae produces extracellular glycerophosphoinositol through phospholipase-mediated turnover of phosphatidylinositol and transports glycerophosphoinositol into the cell upon nutrient limitation. A screening identified the RAS GTPase-activating proteins Ira1 and Ira2 as required for utilization of glycerophosphoinositol as the sole phosphate source, but the RAS/cyclic AMP pathway does not appear to be involved in the growth phenotype. Ira1 and Ira2 affect both the production and transport of glycerophosphoinositol.Membrane phospholipids are continually synthesized and degraded as cells grow and respond to environmental conditions. A major pathway of phosphatidylinositol (PI) turnover in Saccharomyces cerevisiae is its deacylation to produce extracellular glycerophosphoinositol (GroPIns) (3). Plb3, an enzyme with phospholipase B (PLB)/lysophospholipase activity, is thought to be primarily responsible for the production of extracellular GroPIns, with Plb1 playing a lesser role (11, 12, 13). GroPIns is transported into the cell by the Git1 permease (17). GIT1 expression is upregulated by phosphate limitation and inositol limitation. In fact, GroPIns can act as the cell''s sole source of both inositol (17) and phosphate (1).A screening for gene products involved in the process by which GroPIns enters the cellular metabolism identified Ira1 and Ira2, yeast homologs of the mammalian protein neurofibromin. Alterations in NF1, the gene encoding neurofibromin, are associated with the pathogenesis of neurofibromatosis type 1, an autosomal dominant genetic disease (4, 5, 25). Ira1 and Ira2 and neurofibromin function as RAS GTPase-activating proteins (RAS GAPs). S. cerevisiae Ras1 and Ras2 activate adenylate cyclase to modulate cyclic AMP (cAMP) levels. The binding of cAMP to the regulatory subunits of protein kinase A (Bcy1) results in dissociation and activation of the catalytic subunits (Tpk1 to Tpk3). Ira1 and Ira2 inactivate RAS and thereby downregulate the pathway (18, 19). Hydrolysis of cAMP by the phosphodiesterases encoded by PDE1 and PDE2 also downregulate the pathway (7, 20, 23). The RAS/cAMP pathway responds to nutrient signals to modulate fundamental cellular processes, including stress resistance, metabolism, and cell proliferation (7, 20, 21).  相似文献   

15.
16.
Ras leads an important signaling pathway that is deregulated in neurofibromatosis type 1 and malignant peripheral nerve sheath tumor (MPNST). In this study, we show that overactivation of Ras and many of its downstream effectors occurred in only a fraction of MPNST cell lines. RalA, however, was overactivated in all MPNST cells and tumor samples compared to nontransformed Schwann cells. Silencing Ral or inhibiting it with a dominant-negative Ral (Ral S28N) caused a significant reduction in proliferation, invasiveness, and in vivo tumorigenicity of MPNST cells. Silencing Ral also reduced the expression of epithelial mesenchymal transition markers. Expression of the NF1-GTPase-related domain (NF1-GRD) diminished the levels of Ral activation, implicating a role for neurofibromin in regulating RalA activation. NF1-GRD treatment caused a significant decrease in proliferation, invasiveness, and cell cycle progression, but cell death increased. We propose Ral overactivation as a novel cell signaling abnormality in MPNST that leads to important biological outcomes with translational ramifications.The Ras family of guanine-nucleotide bound proteins exerts a fundamental role in cell biology and constitutes an important area of cancer research due to its significant involvement in the development and progression of malignancies (8, 10, 17, 18, 32). Ras-like (Ral) proteins are crucial members of this family and have been shown to play a pivotal role in human tumors (7, 28, 41, 66, 70). Because Ral guanine nucleotide exchange factors (Ral-GEFs) are direct effectors of Ras, the Ral signaling pathway has been traditionally considered a Ras-effector pathway. Activation of Ras (in resemblance to Ral) is regulated by two classes of proteins: Ras-GEFs (e.g., SOS) and Ras- GTPase activating proteins (Ras-GAPs such as neurofibromin). The latter induces hydrolysis of Ras from the active (GTP) form to the inactive (GDP) form (13). Ral-GEFs include two main groups: the proteins that are stimulated by Ras because of their carboxy-terminal Ras binding domain (RalGDS, RGL1, and RGL2) and the proteins that are activated by substrates of PI3K through a pleckstrin homology domain on their C-terminal (RALGPS1 and RALGPS2) (19). Although highly similar to Ras, Ral proteins (RalA and RalB) involve a series of distinctly different effectors that influence gene expression and translation through interaction with ZO-1-associated nucleic acid binding protein (ZONAB) and RalA binding protein 1 (RalBP1) (11, 23, 33). RalB directly interacts with the SEC5 subunit of exocyst to facilitate the host defense response (48, 58).In addition to overactivation of GEFs, inactivation of GAPs is another mechanism for overactivation of GTP-bound proteins. The lack of neurofibromin (encoded by NF1 on human chromosome 17q11.2), a Ras-GAP protein, is the main molecular event in neurofibromatosis type 1 (NF-1), an autosomal-dominant human genetic disease occurring in approximately 1 in 2,500 to 3,500 births (22, 27, 42). One of the main tumor-causing effects of inactivating mutations in the tumor suppressor NF1 gene is postulated to be the subsequent activation of Ras (3, 29, 53, 57, 69). With two main functional domains, SEC14 and Ras-GAP, neurofibromin is best known for its Ras-GAP function. Although the yeast SEC14p is shown to be involved in regulating intracellular proteins and lipid trafficking, the function of its homologous domain in neurofibromin is unknown (49, 62). Although neurofibromas are the most common tumors in NF-1, 10% of patients with plexiform develop malignant peripheral nerve sheath tumors (MPNSTs), which are typically high grade and often fatal (21, 34, 65).The molecular events involved in the malignant transformation of benign neurofibromas to MPNST are poorly defined. Usually arising in the third through sixth decades of life, these tumors are composed of tightly packed hyperchromatic spindle-shaped cells with frequent mitotic figures. Inactivation of both copies of the NF1 gene has been demonstrated in benign human neurofibromas and shown to cause tumors in murine models (56). Loss of heterozygosity of NF1 and p53 has frequently been observed in human MPNST (35, 47, 54). Recombinant mouse strains (NP mice), which harbor inactivated Nf1 and p53 alleles (cis-Nf1+/−:p53+/−), demonstrate the cumulative effects of loss of both Nf1 and p53 genes in the etiology of MPNST (14, 68).In the present study, we show that while both Ras activation and activation of a series of its downstream effector pathways are observed in a fraction of MPNST cells, RalA is activated globally in all studied mouse and human MPNST cells and tumor samples. Our results also explain the involvement of this signaling molecule in a series of key biological functions of MPNST cells, as shown in a variety of in vitro assays and an in vivo model of MPNST. Such information may play a role in designing novel therapies for treatment of MPNST or other tumors with overactivation of the Ral pathway.  相似文献   

17.
Bocavirus is a newly classified genus of the family Parvovirinae. Infection with Bocavirus minute virus of canines (MVC) produces a strong cytopathic effect in permissive Walter Reed/3873D (WRD) canine cells. We have systematically characterized the MVC infection-produced cytopathic effect in WRD cells, namely, the cell death and cell cycle arrest, and carefully examined how MVC infection induces the cytopathic effect. We found that MVC infection induces an apoptotic cell death characterized by Bax translocalization to the mitochondrial outer membrane, disruption of the mitochondrial outer membrane potential, and caspase activation. Moreover, we observed that the activation of caspases occurred only when the MVC genome was replicating, suggesting that replication of the MVC genome induces apoptosis. MVC infection also induced a gradual cell cycle arrest from the S phase in early infection to the G2/M phase at a later stage, which was confirmed by the upregulation of cyclin B1 and phosphorylation of cdc2. Cell cycle arrest at the G2/M phase was reproduced by transfection of a nonreplicative NS1 knockout mutant of the MVC infectious clone, as well as by inoculation of UV-irradiated MVC. In contrast with other parvoviruses, only expression of the MVC proteins by transfection did not induce apoptosis or cell cycle arrest. Taken together, our results demonstrate that MVC infection induces a mitochondrion-mediated apoptosis that is dependent on the replication of the viral genome, and the MVC genome per se is able to arrest the cell cycle at the G2/M phase. Our results may shed light on the molecular pathogenesis of Bocavirus infection in general.The Bocavirus genus is newly classified within the subfamily Parvovirinae of the family Parvoviridae (21). The currently known members of the Bocavirus genus include bovine parvovirus type 1 (BPV1) (17), minute virus of canines (MVC) (57), and the recently identified human bocaviruses (HBoV, HBoV2, and HBoV3) (4, 7, 36).MVC was first recovered from canine fecal samples in 1970 (10). The virus causes respiratory disease with breathing difficulty (14, 32, 49) and enteritis with severe diarrhea (11, 39), which often occurs with coinfection with other viruses (39), spontaneous abortion of fetuses, and death of newborn puppies (14, 29). Pathological lesions in fetuses in experimental infections were found in the lymphoid tissue of the lung and small intestine (14). MVC was isolated and grown in the Walter Reed/3873D (WRD) canine cell line (10), which is derived from a subdermoid cyst of an irradiated male dog (10). The full-length 5.4-kb genome of MVC was recently mapped with palindromic termini (60). Under the control of a single P6 promoter, through the mechanism of alternative splicing and alternative polyadenylation, MVC expresses two nonstructural proteins (NS1 and NP1) and two capsid proteins (VP1 and VP2). Like the NS1 proteins of other parvoviruses, the NS1 of MVC is indispensable for genome replication. The NP1 protein, which is unique to the Bocavirus genus, appears to be critical for optimal viral replication, as the NP1 knockout mutant of MVC suffers from severe impairment of replication (60). A severe cytopathic effect during MVC infection of WRD cells has been documented (10, 60).The HBoV genome has been frequently detected worldwide in respiratory specimens from children under 2 years old with acute respiratory illnesses (2, 34, 55). HBoV is associated with acute expiratory wheezing and pneumonia (3, 34, 55) and is commonly detected in association with other respiratory viruses (34, 55). Further studies are necessary, however, to identify potential associations of HBoV infection with clinical symptoms or disease of acute gastroenteritis (7, 36). The full-length sequence of infectious MVC DNA (GenBank accession no. FJ214110) that we have reported shows 52.6% identity to HBoV, while the NS1, NP1, and VP1 proteins are 38.5%, 39.9%, and 43.7% identical to those of HBoV, respectively (60).The cytopathic effect induced during parvovirus infection has been widely documented, e.g., in infections with minute virus of mice (MVM) (13), human parvovirus B19 (B19V) (58), parvovirus H-1 (25, 52), and BPV1 (1). In Bocavirus, cell death during BPV1 infection of embryonic bovine tracheal cells has been shown to be achieved through necrosis, independent of apoptosis (1). B19V-induced cell death of primary erythroid progenitor cells has been shown to be mainly mediated by an apoptotic pathway (58) in which the nonstructural protein 11kDa plays a key role (16). In contrast, the MVM-induced cytopathic effect has been revealed to be mediated by NS1 interference with intracellular casein kinase II (CKII) signaling (22, 44, 45), a nonapoptotic cell death. Oncolytic parvovirus H-1 infections can induce either apoptosis or nonapoptotic cell death, depending on the cell type (25, 40). Therefore, the mechanisms underlying parvovirus infection-induced cell death vary, although NS1 has been widely shown to be involved in both apoptotic and nonapoptotic cell death. The nature of the cytopathic effect during Bocavirus MVC infection has not been studied.Parvovirus replication requires infected cells at the S phase. Infection with parvovirus has been revealed to accompany a cell cycle perturbation that mostly leads to an arrest in the S/G2 phase or the G2/M phase during infection (30, 33, 42, 47, 65). MVM NS1 expression induces an accumulation of sensitive cells in the S/G2 phase (6, 46, 47). Whether MVC infection-induced cell death is accompanied by an alternation of cell cycle progression and whether the viral nonstructural protein is involved in these processes have not been addressed.In this study, we found, in contrast with other members of the family Parvoviridae, expression of both the nonstructural and structural proteins of MVC by transfection did not induce cell death or cell cycle arrest. However, the cytopathic effect induced during MVC infection is a replication-coupled, mitochondrion-mediated and caspase-dependent apoptosis, accompanied with a gradual cell cycle arrest from the S phase to the G2/M phase, which is facilitated by the MVC genome.  相似文献   

18.
The Target Of Rapamycin (TOR) kinase belongs to the highly conserved eukaryotic family of phosphatidylinositol-3-kinase-related kinases (PIKKs). TOR proteins are found at the core of two distinct evolutionarily conserved complexes, TORC1 and TORC2. Disruption of TORC1 or TORC2 results in characteristically dissimilar phenotypes. TORC1 is a major cell growth regulator, while the cellular roles of TORC2 are not well understood. In the fission yeast Schizosaccharomyces pombe, Tor1 is a component of the TORC2 complex, which is particularly required during starvation and various stress conditions. Our genome-wide gene expression analysis of Δtor1 mutants indicates an extensive similarity with chromatin structure mutants. Consistently, TORC2 regulates several chromatin-mediated functions, including gene silencing, telomere length maintenance, and tolerance to DNA damage. These novel cellular roles of TORC2 are rapamycin insensitive. Cells lacking Tor1 are highly sensitive to the DNA-damaging drugs hydroxyurea (HU) and methyl methanesulfonate, similar to mutants of the checkpoint kinase Rad3 (ATR). Unlike Rad3, Tor1 is not required for the cell cycle arrest in the presence of damaged DNA. Instead, Tor1 becomes essential for dephosphorylation and reactivation of the cyclin-dependent kinase Cdc2, thus allowing reentry into mitosis following recovery from DNA replication arrest. Taken together, our data highlight critical roles for TORC2 in chromatin metabolism and in promoting mitotic entry, most notably after recovery from DNA-damaging conditions. These data place TOR proteins in line with other PIKK members, such as ATM and ATR, as guardians of genome stability.The TOR protein kinase is a major cell growth regulator that links cellular growth with cell divisions (18, 42, 64, 65). TOR is an atypical protein kinase conserved from yeast to humans that was isolated as the target of the immunosuppressive and anticancer drug rapamycin (28). TOR proteins can be found in two distinct complexes, known as TORC1 and TORC2 (27, 64). These complexes mediate their distinct cellular functions via phosphorylation and activation of different sets of AGC-like kinases, including mammalian p70S6K, downstream of TORC1, and AKT/protein kinase B (PKB) downstream of TORC2 (18). TORC1 in mammals contains mTOR (Tor1 or Tor2 in Saccharomyces cerevisiae; Tor2 in Schizosaccharomyces pombe) and the Raptor protein (Kog1 in S. cerevisiae; Mip1 in S. pombe). TORC1 in many different eukaryotes plays a central role in the control of growth (mass accumulation) in response to external stimuli, particularly nutrient availability. Disruption of TORC1, either by mutating its components or by rapamycin treatment, can lead to a starvation-like phenotype (64). The cellular roles of TORC2, on the other hand, are less well defined. TORC2 in mammals contains mTOR (Tor2 in S. cerevisiae; Tor1 in S. pombe) together with Rictor (Avo3 in S. cerevisiae; Ste20 in S. pombe) and mSin1 (Avo1 in S. cerevisiae; Sin1 in S. pombe). TORC2 plays a role in regulating the actin cytoskeleton and cell wall integrity pathway in S. cerevisiae (3, 15, 27), a function that is at least partially conserved in human cells (17, 47).Fission yeast contains two TOR homologues, Tor1 and Tor2 (59), which form the TORC2 and TORC1 complexes, respectively (14, 32). Disruption tor2+ (TORC1) mimics nitrogen starvation responses (1, 14, 32, 56, 57, 62), while disruption of tor1+ (TORC2) results in pleiotropic defects, including elongated cells, sensitivity to osmotic and oxidative stress, inability to execute developmental processes in response to nutrient depletion, and a decrease in amino acid uptake (16, 22, 59). Tor1 regulates cell survival under stress conditions and starvation responses via the AGC protein kinase Gad8, a putative homologue of mammalian AKT/PKB (16).In budding yeast and mammalian cells, TORC1 mediates the rapamycin-sensitive signaling branch while TORC2 is far less sensitive to inhibition by this drug (27, 48). Curiously, rapamycin does not inhibit growth of S. pombe cells but partially inhibits sexual development and amino acid uptake (60-62). Inhibition of amino acid uptake is likely a result of inhibiting Tor1 (61, 62). Accordingly, a tor1 rapamycin-defective allele (tor1S1834E) confers rapamycin resistance to strains that are dependent on amino acid uptake for their growth (61). Yet rapamycin also induces a response similar to that for a shift from rich to poor nitrogen conditions, an effect that may involve inhibition of both Tor1 and Tor2 (41).While other members of the phosphatidylinositol-3-kinase-related kinase (PIKK) family of proteins, such as ATM and ATR, have been shown to play central roles in the DNA damage response, little is known about roles that TOR proteins might play in such processes. Recently it was shown that the rapamycin-sensitive TORC1 complex participates in regulating cell survival under DNA-damaging conditions (24, 42, 49). Currently, no such role has been attributed to TORC2.Here we show that Tor1 (TORC2) is critical for cell survival under DNA-damaging conditions, gene silencing at heterochromatic regions, and telomere length maintenance and for regulation of cell cycle progression. Since the TOR complexes are highly conserved in evolution, this novel TORC2 function may also be conserved in other organisms.  相似文献   

19.
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   

20.
Porcine circovirus type 1 (PCV1), originally isolated as a contaminant of PK-15 cells, is nonpathogenic, whereas porcine circovirus type 2 (PCV2) causes an economically important disease in pigs. To determine the factors affecting virus replication, we constructed chimeric viruses by swapping open reading frame 1 (ORF1) (rep) or the origin of replication (Ori) between PCV1 and PCV2 and compared the replication efficiencies of the chimeric viruses in PK-15 cells. The results showed that the replication factors of PCV1 and PCV2 are fully exchangeable and, most importantly, that both the Ori and rep of PCV1 enhance the virus replication efficiencies of the chimeric viruses with the PCV2 backbone.Porcine circovirus (PCV) is a single-stranded DNA virus in the family Circoviridae (34). Type 1 PCV (PCV1) was discovered in 1974 as a contaminant of porcine kidney cell line PK-15 and is nonpathogenic in pigs (31-33). Type 2 PCV (PCV2) was discovered in piglets with postweaning multisystemic wasting syndrome (PMWS) in the mid-1990s and causes porcine circovirus-associated disease (PCVAD) (1, 9, 10, 25). PCV1 and PCV2 have similar genomic organizations, with two major ambisense open reading frames (ORFs) (16). ORF1 (rep) encodes two viral replication-associated proteins, Rep and Rep′, by differential splicing (4, 6, 21, 22). The Rep and Rep′ proteins bind to specific sequences within the origin of replication (Ori) located in the intergenic region, and both are responsible for viral replication (5, 7, 8, 21, 23, 28, 29). ORF2 (cap) encodes the immunogenic capsid protein (Cap) (26). PCV1 and PCV2 share approximately 80%, 82%, and 62% nucleotide sequence identity in the Ori, rep, and cap, respectively (19).In vitro studies using a reporter gene-based assay system showed that the replication factors of PCV1 and PCV2 are functionally interchangeable (2-6, 22), although this finding has not yet been validated in a live infectious-virus system. We have previously shown that chimeras of PCV in which cap has been exchanged between PCV1 and PCV2 are infectious both in vitro and in vivo (15), and an inactivated vaccine based on the PCV1-PCV2 cap (PCV1-cap2) chimera is used in the vaccination program against PCVAD (13, 15, 18, 27).PCV1 replicates more efficiently than PCV2 in PK-15 cells (14, 15); thus, we hypothesized that the Ori or rep is directly responsible for the differences in replication efficiencies. The objectives of this study were to demonstrate that the Ori and rep are interchangeable between PCV1 and PCV2 in a live-virus system and to determine the effects of swapped heterologous replication factors on virus replication efficiency in vitro.  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