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 46 毫秒
1.
The mammalian target of rapamycin (mTOR) is a conserved Ser/Thr kinase that forms two functionally distinct complexes important for nutrient and growth factor signaling. While mTOR complex 1 (mTORC1) regulates mRNA translation and ribosome biogenesis, mTORC2 plays an important role in the phosphorylation and subsequent activation of Akt. Interestingly, mTORC1 negatively regulates Akt activation, but whether mTORC1 signaling directly targets mTORC2 remains unknown. Here we show that growth factors promote the phosphorylation of Rictor (rapamycin-insensitive companion of mTOR), an essential subunit of mTORC2. We found that Rictor phosphorylation requires mTORC1 activity and, more specifically, the p70 ribosomal S6 kinase 1 (S6K1). We identified several phosphorylation sites in Rictor and found that Thr1135 is directly phosphorylated by S6K1 in vitro and in vivo, in a rapamycin-sensitive manner. Phosphorylation of Rictor on Thr1135 did not affect mTORC2 assembly, kinase activity, or cellular localization. However, cells expressing a Rictor T1135A mutant were found to have increased mTORC2-dependent phosphorylation of Akt. In addition, phosphorylation of the Akt substrates FoxO1/3a and glycogen synthase kinase 3α/β (GSK3α/β) was found to be increased in these cells, indicating that S6K1-mediated phosphorylation of Rictor inhibits mTORC2 and Akt signaling. Together, our results uncover a new regulatory link between the two mTOR complexes, whereby Rictor integrates mTORC1-dependent signaling.The mammalian target of rapamycin (mTOR) is an evolutionarily conserved phosphatidylinositol 3-kinase (PI3K)-related Ser/Thr kinase that integrates signals from nutrients, energy sufficiency, and growth factors to regulate cell growth as well as organ and body size in a variety of organisms (reviewed in references 4, 38, 49, and 77). mTOR was discovered as the molecular target of rapamycin, an antifungal agent used clinically as an immunosuppressant and more recently as an anticancer drug (5, 20). Recent evidence indicates that deregulation of the mTOR pathway occurs in a majority of human cancers (12, 18, 25, 46), suggesting that rapamycin analogs may be potent antineoplastic therapeutic agents.mTOR forms two distinct multiprotein complexes, the rapamycin-sensitive and -insensitive mTOR complexes 1 and 2 (mTORC1 and mTORC2), respectively (6, 47). In cells, rapamycin interacts with FKBP12 and targets the FKBP12-rapamycin binding (FRB) domain of mTORC1, thereby inhibiting some of its function (13, 40, 66). mTORC1 is comprised of the mTOR catalytic subunit and four associated proteins, Raptor (regulatory associated protein of mTOR), mLST8 (mammalian lethal with sec13 protein 8), PRAS40 (proline-rich Akt substrate of 40 kDa), and Deptor (28, 43, 44, 47, 59, 73, 74). The small GTPase Rheb (Ras homolog enriched in brain) is a key upstream activator of mTORC1 that is negatively regulated by the tuberous sclerosis complex 1 (TSC1)/TSC2 GTPase-activating protein complex (reviewed in reference 35). mTORC1 is activated by PI3K and Ras signaling through direct phosphorylation and inactivation of TSC2 by Akt, extracellular signal-regulated kinase (ERK), and p90 ribosomal protein S6 kinase (RSK) (11, 37, 48, 53, 63). mTORC1 activity is also regulated at the level of Raptor. Whereas low cellular energy levels negatively regulate mTORC1 activity through phosphorylation of Raptor by AMP-activated protein kinase (AMPK) (27), growth signaling pathways activating the Ras/ERK pathway positively regulate mTORC1 activity through direct phosphorylation of Raptor by RSK (10). More recent evidence has also shown that mTOR itself positively regulates mTORC1 activity by directly phosphorylating Raptor at proline-directed sites (20a, 75). Countertransport of amino acids (55) and amino acid signaling through the Rag GTPases were also shown to regulate mTORC1 activity (45, 65). When activated, mTORC1 phosphorylates two main regulators of mRNA translation and ribosome biogenesis, the AGC (protein kinase A, G, and C) family kinase p70 ribosomal S6 kinase 1 (S6K1) and eukaryotic initiation factor 4E-binding protein 1 (4E-BP1), and thus stimulates protein synthesis and cellular growth (50, 60).The second mTOR complex, mTORC2, is comprised of mTOR, Rictor (rapamycin-insensitive companion of mTOR), mSin1 (mammalian stress-activated mitogen-activated protein kinase-interacting protein 1), mLST8, PRR5 (proline-rich region 5), and Deptor (21, 39, 58, 59, 66, 76, 79). Rapamycin does not directly target and inhibit mTORC2, but long-term treatment with this drug was shown to correlate with mTORC2 disassembly and cytoplasmic accumulation of Rictor (21, 39, 62, 79). Whereas mTORC1 regulates hydrophobic motif phosphorylation of S6K1, mTORC2 has been shown to phosphorylate other members of the AGC family of kinases. Biochemical and genetic evidence has demonstrated that mTORC2 phosphorylates Akt at Ser473 (26, 39, 68, 70), thereby contributing to growth factor-mediated Akt activation (6, 7, 52). Deletion or knockdown of the mTORC2 components mTOR, Rictor, mSin1, and mLST8 has a dramatic effect on mTORC2 assembly and Akt phosphorylation at Ser473 (26, 39, 79). mTORC2 was also shown to regulate protein kinase Cα (PKCα) (26, 66) and, more recently, serum- and glucocorticoid-induced protein kinase 1 (SGK1) (4, 22). Recent evidence implicates mTORC2 in the regulation of Akt and PKCα phosphorylation at their turn motifs (19, 36), but whether mTOR directly phosphorylates these sites remains a subject of debate (4).Activation of mTORC1 has been shown to negatively regulate Akt phosphorylation in response to insulin or insulin-like growth factor 1 (IGF1) (reviewed in references 30 and 51). This negative regulation is particularly evident in cell culture models with defects in the TSC1/TSC2 complex, where mTORC1 and S6K1 are constitutively activated. Phosphorylation of insulin receptor substrate-1 (IRS-1) by mTORC1 (72) and its downstream target S6K1 has been shown to decrease its stability and lead to an inability of insulin or IGF1 to activate PI3K and Akt (29, 69). Although the mechanism is unknown, platelet-derived growth factor receptor β (PDGF-Rβ) has been found to be downregulated in TSC1- and TSC2-deficient murine embryonic fibroblasts (MEFs), contributing to a reduction of PI3K signaling (80). Interestingly, inhibition of Akt phosphorylation by mTORC1 has also been observed in the presence of growth factors other than IGF-1, insulin, or PDGF, suggesting that there are other mechanisms by which mTORC1 activation restricts Akt activity in cells (reviewed in references 6 and 31). Recent evidence demonstrates that rapamycin treatment causes a significant increase in Rictor electrophoretic mobility (2, 62), suggesting that phosphorylation of the mTORC2 subunit Rictor may be regulated by mTORC1 or downstream protein kinases.Herein, we demonstrate that Rictor is phosphorylated by S6K1 in response to mTORC1 activation. We demonstrate that Thr1135 is directly phosphorylated by S6K1 and found that a Rictor mutant lacking this phosphorylation site increases Akt phosphorylation induced by growth factor stimulation. Cells expressing the Rictor T1135A mutant were found to have increased Akt signaling to its substrates compared to Rictor wild-type- and T1135D mutant-expressing cells. Together, our results suggest that Rictor integrates mTORC1 signaling via its phosphorylation by S6K1, resulting in the inhibition of mTORC2 and Akt signaling.  相似文献   

2.
3.
mTOR, the mammalian target of rapamycin, is a critical node for control of cell growth and survival and has widely been implicated in cancer survival signals. mTOR exists in two complexes: mTORC1 and mTORC2. Phospholipase D (PLD) and its metabolite phosphatidic acid (PA) have been implicated in the regulation of mTOR; however, their role has been controversial. We report here that suppression of PLD prevents phosphorylation of the mTORC1 substrate S6 kinase (S6K) at Thr389 and the mTORC2 substrate Akt at Ser473. Suppression of PLD also blocked insulin-stimulated Akt phosphorylation at Ser473 and the mTORC2-dependent phosphorylation of PRAS40. Importantly, PA was required for the association of mTOR with Raptor to form mTORC1 and that of mTOR with Rictor to form mTORC2. The effect of PA was competitive with rapamycin—with much higher concentrations of rapamycin needed to compete with the PA-mTORC2 interaction than with PA-mTORC1. Suppressing PA production substantially increased the sensitivity of mTORC2 to rapamycin. Data provided here demonstrate a PA requirement for the stabilization of both mTORC1 and mTORC2 complexes and reveal a mechanism for the inhibitory effect of rapamycin on mTOR. This study also suggests that by suppressing PLD activity, mTORC2 could be targeted therapeutically with rapamycin.It has become apparent during the past decade that a critical aspect of tumor progression is the suppression of default apoptotic programs that constitute what is likely the most important protection against cancer. Cellular signals that suppress apoptosis have come to be known as “survival signals.” A common node for survival signals is mTOR, the mammalian target of rapamycin (5, 13, 14, 25). mTOR exists in two distinct complexes, mTORC1 and mTORC2 (21), that differ in their subunit composition and sensitivity to rapamycin. mTORC1 consists of a complex that includes mTOR and a protein known as Raptor (regulatory associated protein of mTOR), whereas mTORC2 consists of a complex that includes mTOR and a protein known as Rictor (rapamycin-insensitive companion of mTOR) (13, 14). There are also mTORC2 complexes that can be distinguished by association with different isoforms of mSin1 (9). While much is known about the regulation of mTORC1 (21), very little is known about the regulation of mTORC2.mTORC1 is highly sensitive to rapamycin, whereas mTORC2 is relatively insensitive to rapamycin (21). However, it was recently reported that long-term exposure to rapamycin prevented the formation of mTORC2 complexes and blocked the phosphorylation of the mTORC2 substrate Akt at Ser473 (24, 38). Rapamycin, in association with FK506 binding protein 12 (FKBP12), has been reported to interact with mTOR in a manner that is competitive with phosphatidic acid (PA), the metabolic product of phospholipase D (PLD) (2, 4). PLD, like mTOR, has been implicated in survival signals in several human cancer cell lines (1, 10, 11, 27, 32, 39). Since rapamycin-FKBP12 competes with PA for binding to mTOR, the sensitivity of mTORC2 complex formation to rapamycin suggests that PA facilitates the assembly of mTORC2—and ultimately the activation of mTORC2. We report here that the assembly of both mTORC1 and mTORC2 complexes is dependent upon PLD and its metabolite PA. The study also provides mechanistic insight into how rapamycin impacts on mTOR-mediated signals and how PLD regulates mTOR by facilitating the formation of mTOR complexes.  相似文献   

4.
The mammalian target of rapamycin (mTOR) interacts with raptor to form the protein complex mTORC1 (mTOR complex 1), which plays a central role in the regulation of cell growth in response to environmental cues. Given that glucose is a primary fuel source and a biosynthetic precursor, how mTORC1 signaling is coordinated with glucose metabolism has been an important question. Here, we found that the glycolytic enzyme glyceraldehyde-3-phosphate dehydrogenase (GAPDH) binds Rheb and inhibits mTORC1 signaling. Under low-glucose conditions, GAPDH prevents Rheb from binding to mTOR and thereby inhibits mTORC1 signaling. High glycolytic flux suppresses the interaction between GAPDH and Rheb and thus allows Rheb to activate mTORC1. Silencing of GAPDH or blocking of the Rheb-GAPDH interaction desensitizes mTORC1 signaling to changes in the level of glucose. The GAPDH-dependent regulation of mTORC1 in response to glucose availability occurred even in TSC1-deficient cells and AMPK-silenced cells, supporting the idea that the GAPDH-Rheb pathway functions independently of the AMPK axis. Furthermore, we show that glyceraldehyde-3-phosphate, a glycolytic intermediate that binds GAPDH, destabilizes the Rheb-GAPDH interaction even under low-glucose conditions, explaining how high-glucose flux suppresses the interaction and activates mTORC1 signaling. Taken together, our results suggest that the glycolytic flux regulates mTOR''s access to Rheb by regulating the Rheb-GAPDH interaction, thereby allowing mTORC1 to coordinate cell growth with glucose availability.The mTOR complex 1 (mTORC1) signal transduction pathway acts as a central controller of cell growth in mammals (20, 23, 29). mTORC1 integrates a wide range of intracellular and extracellular signals, including insulin, availability of nutrients (glucose and amino acids), cellular energy status, and hypoxia, to regulate protein synthesis and cell growth (11, 12, 17, 36, 46). Many of these environmental cues are integrated into tuberous sclerosis complex (TSC1-TSC2), the major upstream regulator of mTORC1. In response to the absence of insulin and to the low-energy status of cells, the TSC1-TSC2 complex stimulates the GTPase function of Rheb, a small GTPase that acts as a proximal key activator of mTORC1, which leads to the inhibition of Rheb-mediated mTORC1 activation. In contrast, inactivation of the TSC1-TSC2 complex results in the accumulation of GTP-bound Rheb and thus activation of mTORC1 (3, 13, 21, 27, 32, 39). For this reason, both the loss of TSC proteins and the overexpression of Rheb cause hyperactivation of mTORC1 signaling, which is frequently observed in many common human cancers (2, 5, 19, 25, 33). Therefore, a tight regulation of Rheb activity is critical for the proper operation of the mTORC1 pathway in response to environmental cues.Rheb is an atypical member of the Ras superfamily of GTPases (1, 10, 47). As with other small GTPases, the activity of Rheb is regulated by its guanine nucleotide binding status. However, the negative control of GTP-bound Rheb by the TSC1-TSC2 complex has only recently been investigated, and the regulation of the nucleotide binding status of Rheb is not fully understood. A recent study proposed that translationally controlled tumor protein may function as a guanine nucleotide exchange factor for Rheb that causes the accumulation of GTP-bound Rheb (18). GTP-bound Rheb is essential for activating mTOR kinase (21, 28, 38). However, the interaction between Rheb and mTOR does not depend on the GTP binding status of Rheb (30), raising questions regarding the mechanism by which Rheb activates mTORC1. Recently, FKBP38 (immunophilin FK506-binding protein, 38 kDa) was found to be a direct binding partner of Rheb and an inhibitor of mTORC1 (4). GTP-bound Rheb binds FKBP38 and releases FKBP38 from mTORC1, resulting in activation of the mTORC1 pathway. However, there have been conflicting results regarding the effects of nutrient availability on Rheb activity (31, 37, 42, 50) and the effect of these newly identified regulators of Rheb function (44, 45). Thus, the precise molecular mechanisms underlying Rheb regulation and Rheb-mediated mTORC1 activation have remained unclear.In this study, we identified glyceraldehyde-3-phosphate (Gly-3-P) dehydrogenase (GAPDH) as a novel Rheb binding protein and a negative regulator of Rheb. We found that the interaction between GAPDH and Rheb is induced when the glycolytic flux is suppressed under low-glucose conditions to inhibit mTORC1. Here, we provide a molecular mechanism underlying the cross talk between the glycolytic flux and the mTORC1 signaling.  相似文献   

5.
6.
7.
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   

8.
9.
mTORC1 contains multiple proteins and plays a central role in cell growth and metabolism. Raptor (regulatory-associated protein of mammalian target of rapamycin (mTOR)), a constitutively binding protein of mTORC1, is essential for mTORC1 activity and critical for the regulation of mTORC1 activity in response to insulin signaling and nutrient and energy sufficiency. Herein we demonstrate that mTOR phosphorylates raptor in vitro and in vivo. The phosphorylated residues were identified by using phosphopeptide mapping and mutagenesis. The phosphorylation of raptor is stimulated by insulin and inhibited by rapamycin. Importantly, the site-directed mutation of raptor at one phosphorylation site, Ser863, reduced mTORC1 activity both in vitro and in vivo. Moreover, the Ser863 mutant prevented small GTP-binding protein Rheb from enhancing the phosphorylation of S6 kinase (S6K) in cells. Therefore, our findings indicate that mTOR-mediated raptor phosphorylation plays an important role on activation of mTORC1.Mammalian target of rapamycin (mTOR)2 has been shown to function as a critical controller in cellular growth, survival, metabolism, and development (1). mTOR, a highly conserved Ser-Thr phosphatidylinositol 3-kinase-related protein kinase, structurally forms two distinct complexes, mTOR complex 1 (mTORC1) and mTOR complex 2 (mTORC2), each of which catalyzes the phosphorylation of different substrates (1). The best characterized substrates for mTORC1 are eIF4E-binding protein (4E-BP, also known as PHAS) and p70 S6 kinase (S6K) (1), whereas mTORC2 phosphorylates the hydrophobic and turn motifs of protein kinase B (Akt/protein kinase B) (2) and protein kinase C (3, 4). mTORC1 constitutively consists of mTOR, raptor, and mLst8/GβL (1), whereas the proline-rich Akt substrate of 40 kDa (PRAS40) is a regulatory component of mTORC1 that disassociates after growth factor stimulation (5, 6). Raptor is essential for mTORC1 activity by providing a substrate binding function (7) but also plays a regulatory role on mTORC1 with stimuli of growth factors and nutrients (8). In response to insulin, raptor binding to substrates is elevated through the release of the competitive inhibitor PRAS40 from mTORC1 (9, 10) because PRAS40 and the substrates of mTORC1 (4E-BP and S6K) appear to bind raptor through a consensus sequence, the TOR signaling (TOS) motif (1014). In response to amino acid sufficiency, raptor directly interacts with a heterodimer of Rag GTPases and promotes mTORC1 localization to the Rheb-containing vesicular compartment (15).mTORC1 integrates signaling pathways from growth factors, nutrients, energy, and stress, all of which generally converge on the tuberous sclerosis complex (TSC1-TSC2) through the phosphorylation of TSC2 (1). Growth factors inhibit the GTPase-activating protein activity of TSC2 toward the small GTPase Rheb via the PI3K/Akt pathway (16, 17), whereas energy depletion activates TSC2 GTPase-activating protein activity by stimulating AMP-activated protein kinase (AMPK) (18). Rheb binds directly to mTOR, albeit with very low affinity (19), and upon charging with GTP, Rheb functions as an mTORC1 activator (6). mTORC1 complexes isolated from growth factor-stimulated cells show increased kinase activity yet do not contain detectable levels of associated Rheb. Therefore, how Rheb-GTP binding to mTOR leads to an increase in mTORC1 activity toward substrates, and what the role of raptor is in this activation is currently unknown. More recently, the AMPK and p90 ribosomal S6 kinase (RSK) have been reported to directly phosphorylate raptor and regulate mTORC1 activity. The phosphorylation of raptor directly by AMPK reduced mTORC1 activity, suggesting an alternative regulation mechanism independent of TSC2 in response to energy supply (20). RSK-mediated raptor phosphorylation enhances mTORC1 activity and provides a mechanism whereby stress may activate mTORC1 independent of the PI3K/Akt pathway (21). Therefore, the phosphorylation status of raptor can be critical for the regulation of mTORC1 activity.In this study, we investigated phosphorylation sites in raptor catalyzed by mTOR. Using two-dimensional phosphopeptide mapping, we found that Ser863 and Ser859 in raptor were phosphorylated by mTOR both in vivo and in vitro. mTORC1 activity in vitro and in vivo is associated with the phosphorylation of Ser863 in raptor.  相似文献   

10.
The mTOR signaling complex integrates signals from growth factors and nutrient availability to control cell growth and proliferation, in part through effects on the protein-synthetic machinery. Protein synthesis rates fluctuate throughout the cell cycle but diminish significantly during the G2/M transition. The fate of the mTOR complex and its role in coordinating cell growth and proliferation signals with protein synthesis during mitosis remain unknown. Here we demonstrate that the mTOR complex 1 (mTORC1) pathway, which stimulates protein synthesis, is actually hyperactive during mitosis despite decreased protein synthesis and reduced activity of mTORC1 upstream activators. We describe previously unknown G2/M-specific phosphorylation of a component of mTORC1, the protein raptor, and demonstrate that mitotic raptor phosphorylation alters mTORC1 function during mitosis. Phosphopeptide mapping and mutational analysis demonstrate that mitotic phosphorylation of raptor facilitates cell cycle transit through G2/M. Phosphorylation-deficient mutants of raptor cause cells to delay in G2/M, whereas depletion of raptor causes cells to accumulate in G1. We identify cyclin-dependent kinase 1 (cdk1 [cdc2]) and glycogen synthase kinase 3 (GSK3) pathways as two probable mitosis-regulated protein kinase pathways involved in mitosis-specific raptor phosphorylation and altered mTORC1 activity. In addition, mitotic raptor promotes translation by internal ribosome entry sites (IRES) on mRNA during mitosis and is demonstrated to be associated with rapamycin resistance. These data suggest that this pathway may play a role in increased IRES-dependent mRNA translation during mitosis and in rapamycin insensitivity.Cell growth and cell division are tightly coordinated processes required for cells to remain equal in size after division. In unicellular organisms, cell growth and proliferation are coordinated by nutrient availability, whereas their multicellular counterparts must also respond to growth factor input. Both processes lead to organismal growth as well as to increased cell number and cell mass. Cell growth and cell proliferation are also linked via the mTOR signaling pathway (16, 17). The mTOR kinase forms a distinct signaling complex (mTORC1) that participates in the coordination of nutrient and growth factor signaling. mTORC1 is composed of the kinase mTOR, the adaptor protein raptor, and the regulatory protein LST8 (25, 33, 34, 72).Accumulation of cellular proteins leads to cell growth and cell division. However, cell growth occurs only during certain phases of the cell cycle, necessitating that protein synthesis rates oscillate during cell cycling (40). In addition, in quiescent cells in G0, protein synthesis rates are significantly reduced, whereas a select group of mRNAs maintain active translation (20, 68). During the G1 phase, overall protein synthesis rates increase through S phase to allow cells to grow and enter another round of cell division while maintaining cell size (2, 3, 42, 45). As with G0, entrance into mitosis (G2/M phase) results in a global downregulation by as much as 60 to 80% of cap-dependent mRNA translation in primary, immortalized, and some transformed cells (5, 14, 29).Studies report several possible mechanisms for inhibition of protein synthesis during mitosis. Translation initiation requires the formation of an initiation factor complex known as eukaryotic translation initiation factor 4F (eIF4F), which consists of cap binding protein eIF4E, molecular scaffold protein eIF4G, and RNA helicase eIF4A. Together, they recruit ribosomes to mRNAs via bridging interactions between the 7-methyl-GTP (m7GTP) 5′ cap and the small 40S ribosomal subunit. Downregulation of protein synthesis during G2/M was first ascribed to hypophosphorylation of eIF4E and the eIF4E binding proteins (4E-BPs) (5, 46). 4E-BPs are activated by hypophosphorylation, which allows them to bind and sequester eIF4E, preventing it from binding eIF4G and thereby blocking cap-dependent mRNA translation. More recently, several studies suggest that 4E-BP1, the major 4E-BP and a key target of mTORC1, is actually hyperphosphorylated (inactivated) during mitosis (26, 49). It is puzzling, then, that the phosphatidylinositol 3-kinase (PI3K)/AKT network and AKT itself (which modulate mTORC1 activity) are reportedly inactivated during late mitosis (1, 9, 22). In addition, phosphorylation of another mTORC1 target, ribosomal S6 kinase 1 (S6K1), and its activity are actually highest during G2/M phase, consistent with elevated mTORC1 activity during mitosis (6).In this study we show that, despite repression of AKT and other activators of mTORC1 activity in mitosis, mTORC1 remains active and phosphorylates 4E-BP1 and S6K1 during G2/M. We describe the multisite phosphorylation of raptor during mitosis, and we identify seven mitosis-specific raptor phosphorylation sites. By developing phosphomimetic and phosphorylation-deficient mutants of raptor, we show that hyperphosphorylated raptor promotes cell cycle transit through G2/M, whereas hypophosphorylated raptor promotes transit through G1. Raptor phosphorylation is shown to involve kinase pathways that are known to be active during mitosis, including cyclin-dependent kinase 1 (cdk1 [cdc2]) and glycogen synthase kinase 3 (GSK3) pathways that are also upregulated in certain human cancers, including breast cancers. These and other findings disclose a novel regulatory network for mTORC1 that is active during mitosis, important for G2/M progression and increased internal ribosome entry site (IRES)-dependent translation during mitosis, and indirectly associated with rapamycin resistance.  相似文献   

11.
Rheb G-protein plays critical roles in the TSC/Rheb/mTOR signaling pathway by activating mTORC1. The activation of mTORC1 by Rheb can be faithfully reproduced in vitro by using mTORC1 immunoprecipitated by the use of anti-raptor antibody from mammalian cells starved for nutrients. The low in vitro kinase activity against 4E-BP1 of this mTORC1 preparation is dramatically increased by the addition of recombinant Rheb. On the other hand, the addition of Rheb does not activate mTORC2 immunoprecipitated from mammalian cells by the use of anti-rictor antibody. The activation of mTORC1 is specific to Rheb, because other G-proteins such as KRas, RalA/B, and Cdc42 did not activate mTORC1. Both Rheb1 and Rheb2 activate mTORC1. In addition, the activation is dependent on the presence of bound GTP. We also find that the effector domain of Rheb is required for the mTORC1 activation. FKBP38, a recently proposed mediator of Rheb action, appears not to be involved in the Rheb-dependent activation of mTORC1 in vitro, because the preparation of mTORC1 that is devoid of FKBP38 is still activated by Rheb. The addition of Rheb results in a significant increase of binding of the substrate protein 4E-BP1 to mTORC1. PRAS40, a TOR signaling (TOS) motif-containing protein that competes with the binding of 4EBP1 to mTORC1, inhibits Rheb-induced activation of mTORC1. A preparation of mTORC1 that is devoid of raptor is not activated by Rheb. Rheb does not induce autophosphorylation of mTOR. These results suggest that Rheb induces alteration in the binding of 4E-BP1 with mTORC1 to regulate mTORC1 activation.Rheb defines a unique member of the Ras superfamily G-proteins (1). We have shown that Rheb proteins are conserved and are found from yeast to human (2). Although yeast and fruit fly have one Rheb, mouse and human have two Rheb proteins termed Rheb1 (or simply Rheb) and Rheb2 (RhebL1) (2). Structurally, these proteins contain G1-G5 boxes, short stretches of amino acids that define the function of the Ras superfamily G-proteins including guanine nucleotide binding (1, 3, 4). Rheb proteins have a conserved arginine at residue 15 that corresponds to residue 12 of Ras (1). The effector domain required for the binding with downstream effectors encompasses the G2 box and its adjacent sequences (1, 5). Structural analysis by x-ray crystallography further shows that the effector domain is exposed to solvent, is located close to the phosphates of GTP especially at residues 35–38, and undergoes conformational change during GTP/GDP exchange (6). In addition, all Rheb proteins end with the CAAX (C is cysteine, A is an aliphatic amino acid, and X is the C-terminal amino acid) motif that signals farnesylation. In fact, we as well as others have shown that these proteins are farnesylated (79).Rheb plays critical roles in the TSC/Rheb/mTOR signaling, a signaling pathway that plays central roles in regulating protein synthesis and growth in response to nutrient, energy, and growth conditions (1014). Rheb is down-regulated by a TSC1·TSC2 complex that acts as a GTPase-activating protein for Rheb (1519). Recent studies established that the GAP domain of TSC2 defines the functional domain for the down-regulation of Rheb (20). Mutations in the Tsc1 or Tsc2 gene lead to tuberous sclerosis whose symptoms include the appearance of benign tumors called hamartomas at different parts of the body as well as neurological symptoms (21, 22). Overexpression of Rheb results in constitutive activation of mTOR even in the absence of nutrients (15, 16). Two mTOR complexes, mTORC1 and mTORC2, have been identified (23, 24). Whereas mTORC1 is involved in protein synthesis activation mediated by S6K and 4EBP1, mTORC2 is involved in the phosphorylation of Akt in response to insulin. It has been suggested that Rheb is involved in the activation of mTORC1 but not mTORC2 (25).Although Rheb is clearly involved in the activation of mTOR, the mechanism of activation has not been established. We as well as others have suggested a model that involves the interaction of Rheb with the TOR complex (2628). Rheb activation of mTOR kinase activity using immunoprecipitated mTORC1 was reported (29). Rheb has been shown to interact with mTOR (27, 30), and this may involve direct interaction of Rheb with the kinase domain of mTOR (27). However, this Rheb/mTOR interaction is a weak interaction and is not dependent on the presence of GTP bound to Rheb (27, 28). Recently, a different model proposing that FKBP38 (FK506-binding protein 38) mediates the activation of mTORC1 by Rheb was proposed (31, 32). In this model, FKBP38 binds mTOR and negatively regulates mTOR activity, and this negative regulation is blocked by the binding of Rheb to FKBP38. However, recent reports dispute this idea (33).To further characterize Rheb activation of mTOR, we have utilized an in vitro system that reproduces activation of mTORC1 by the addition of recombinant Rheb. We used mTORC1 immunoprecipitated from nutrient-starved cells using anti-raptor antibody and have shown that its kinase activity against 4E-BP1 is dramatically increased by the addition of recombinant Rheb. Importantly, the activation of mTORC1 is specific to Rheb and is dependent on the presence of bound GTP as well as an intact effector domain. FKBP38 is not detected in our preparation and further investigation suggests that FKBP38 is not an essential component for the activation of mTORC1 by Rheb. Our study revealed that Rheb enhances the binding of a substrate 4E-BP1 with mTORC1 rather than increasing the kinase activity of mTOR.  相似文献   

12.
13.
Viral manipulation of the transduction pathways associated with key cellular functions such as actin remodeling, microtubule stabilization, and survival may favor a productive viral infection. Here we show that consistent with the vaccinia virus (VACV) and cowpox virus (CPXV) requirement for cytoskeleton alterations early during the infection cycle, PBK/Akt was phosphorylated at S473 [Akt(S473-P)], a modification associated with the mammalian target of rapamycin complex 2 (mTORC2), which was paralleled by phosphorylation at T308 [Akt(T308-P)] by PI3K/PDK1, which is required for host survival. Notably, while VACV stimulated Akt(S473-P/T308-P) at early (1 h postinfection [p.i.]) and late (24 h p.i.) times during the infective cycle, CPXV stimulated Akt at early times only. Pharmacological and genetic inhibition of PI3K (LY294002) or Akt (Akt-X and a dominant-negative form of Akt-K179M) resulted in a significant decline in virus yield (from 80% to ≥90%). This decline was secondary to the inhibition of late viral gene expression, which in turn led to an arrest of virion morphogenesis at the immature-virion stage of the viral growth cycle. Furthermore, the cleavage of both caspase-3 and poly(ADP-ribose) polymerase and terminal deoxynucleotidyl transferase-mediated deoxyuridine nick end labeling assays confirmed that permissive, spontaneously immortalized cells such as A31 cells and mouse embryonic fibroblasts (MEFs) underwent apoptosis upon orthopoxvirus infection plus LY294002 treatment. Thus, in A31 cells and MEFs, early viral receptor-mediated signals transmitted via the PI3K/Akt pathway are required and precede the expression of viral antiapoptotic genes. Additionally, the inhibition of these signals resulted in the apoptosis of the infected cells and a significant decline in viral titers.The family Poxviridae is a family of large, linear, double-stranded DNA viruses that carry out their entire life cycle within the cytoplasmic compartment of infected cells. Vaccinia virus (VACV) is a prototypical member of the genus Orthopoxvirus, which also includes the closely related cowpox virus (CPXV) (12, 52). The genomes of these viruses are approximately 200 kbp in length, with a coding capacity of approximately 200 genes. The genes involved in virus-host interactions are situated at both ends of the genome and are associated with the evasion of host immune defenses (1). These evasion mechanisms operate mainly extracellularly. For example, the secretion of soluble cytokine and chemokine receptor homologues blocks the receptor recognition by intercepting the cognate cytokine/chemokine in the extracellular environment. This mechanism facilitates viral attachment and entry into cells (1, 70). Therefore, decoy receptors for alpha interferon (IFN-α), IFN-β, IFN-γ, and tumor necrosis factor alpha play an important immunomodulatory role by affecting both the host antiviral and apoptotic responses.To counteract the host proapoptotic response, poxviruses have developed a number of antiapoptotic strategies, including the inhibition of apoptotic signals triggered by the extrinsic pathway (those mediated by death receptors such as tumor necrosis factor and Fas ligand) or the intrinsic pathway (mediated by the mitochondria and triggered upon viral infection) (1, 25, 70, 74). Many studies previously identified viral inhibitors that block specific steps of the intrinsic pathway. These include the VACV-encoded E3L, F1L, and N1L genes and the myxoma virus (MYXV)-encoded M11L gene, which block cytochrome c release (14, 20, 34, 39, 45, 75, 90), and the CPXV-encoded cytokine response modifier gene (CrmA) as well as the VACV-encoded SPI-2 gene, which inhibits both caspase-1 and caspase-8 (25, 58, 61, 74).An emerging body of evidence has also highlighted the pivotal role played by intracellular signaling pathways in Orthopoxvirus biology (18, 48, 92). We and others have shown that poxvirus manipulation of signaling pathways can be virus specific. For example, while both VACV and CPXV stimulate the MEK/extracellular signal-regulated kinase (ERK)/EGR-1 pathway during a substantial length of time of their infective cycle, the pathway is required only for VACV replication, whereas its role in CPXV biology has yet to be identified (71). MYXV, a rabbit-specific poxvirus, also activates the MEK/ERK pathway in a mouse model of poxvirus-host interactions. However, this stimulation led to the expression of IFN-β, which consequently blocked virus replication and possibly explains why MYXV has such a restricted host range (87).Another signaling molecule associated with viral replication is Akt kinase (also known as protein kinase B). The MYXV host range factor M-T5 is able to reprogram the intracellular environment, thereby increasing human tumor cell permissiveness to viral replication, which is directly associated with levels of phosphorylated Akt (88). In addition, M-T5 is functionally replaced by the host phosphatidylinositol 3-kinase (PI3K) enhancer A protein (92).The transmission of intracellular signals mediated by the serine/threonine kinase Akt to downstream molecules in response to diverse stimuli such as growth factors, insulin, and hormones is dependent upon the phosphorylation of serine 473 (S473-P) and threonine 308 (T308-P). This phosphorylation is mediated by mammalian target of rapamycin complex 2 (mTORC2) and phosphoinositide-dependent protein kinase 1 (PDK1), which act as downstream effectors of the PI3K/Akt/mTORC1 pathway (2, 66). PI3Ks are a family of enzymes (classes I to III) that generate lipid second messengers by the phosphorylation of plasma membrane phosphoinositides. Class IA PI3Ks consist of a catalytic subunit (p110, comprising the three isoforms α, β, and δ) and an adaptor/regulatory subunit (p85, comprising the two isoforms α and β) (for a detailed review, see reference 80).The Akt family of proteins is comprised of the three isoforms α, β, and γ, which are composed of an N-terminal pleckstrin homology domain, a central catalytic domain, and a C-terminal hydrophobic domain. Akt is recruited to the plasma membrane through the binding of its pleckstrin homology domain to the phosphatidylinositol 3,4,5-triphosphate (PIP3), which is a product of PI3K that is anchored to the plasma membrane. PDK1 is also recruited to the plasma membrane through interactions with PIP3. As both PDK1 and Akt interact with PIP3, PDK1 colocalizes with Akt and activates it by phosphorylating threonine 308 (T308-P) (2, 66). Following its activation, Akt phosphorylates a number of downstream substrates such as caspase-9, BAD, glycogen synthase kinase 3β (GSK-3β), and FKHR. This leads to the suppression of apoptosis, cell growth, survival, and proliferation (11, 16, 56).Another downstream target of PI3K/Akt is mTOR, a serine/threonine kinase that plays a central role in the regulation of cell growth, proliferation, survival, and protein synthesis (26). mTOR kinase has recently been found to be associated with two functionally distinct complexes in mammalian cells, known as mTORC1 and mTORC2 (63, 66). Although these multiprotein complexes share molecules in common, distinct adaptor proteins are recruited into each complex: regulatory-associated protein of TOR (raptor) is recruited into mTORC1, while rapamycin-insensitive companion of TOR (rictor) is recruited into mTORC2 (33, 64). While mTORC1 controls cell growth and protein translation and has proven to be rapamycin sensitive, mTORC2 regulates the actin cytoskeleton and is assumed to be rapamycin insensitive, even though under conditions of prolonged exposure to the drug, it appears to inhibit mTORC2 assembly (29, 64, 65). Additionally, it has been demonstrated that mTORC2 regulates the activity of Akt through the phosphorylation of S473 (S473-P). S473-P appears to be required for the full activation of Akt, since S473-P has been shown to enhance the subsequent phosphorylation of T308 by PDK1 (66, 67, 94). Moreover, the phosphorylation of both S473 and T308 results in a four- to fivefold increase in Akt activity compared to T308-P by PDK1 alone (66).The PI3K/PDK1/Akt(T308)/mTORC1 pathway regulates vital cellular processes that are important for viral replication and propagation, including cell growth, proliferation, and protein translation. This pathway is particularly important for the replication of DNA viruses, as their replication is cap dependent. However, the Akt signaling pathway can also negatively affect viral replication. The stress response downstream of Akt signaling, including hypoxia and energy and amino acid depletion, inhibits mTORC1 (5, 9, 69). Therefore, DNA viruses must overcome these constraints to translate their mRNAs.Pharmacological disruption of the PI3K/Akt pathway with the specific PI3K inhibitor LY294002 (2-morpholino-8-phenyl-4H-1-benzopyran-4-one) (82) has been reported to not only increase the cleavage of downstream molecules associated with proapoptotic activity [e.g., poly(ADP-ribose) polymerase (PARP) and the executioner caspase-3] (38, 41) but also promote microtubule stabilization, actin filament remodeling/cell migration, and bleb formation/viral infectivity (10, 35, 49, 54, 59).Because the PI3K/Akt and PI3K/Akt/mTOR pathways influence diverse cellular functions and possibly a healthy antiviral response, usurping these pathways could support an increase in viral replication. In support of this, a number of reports have demonstrated that either the PI3K/Akt or the PI3K/Akt/mTOR pathway plays a role in the replication of many viruses including flavivirus (38), hepatitis C virus (27), human immunodeficiency virus type 1 (93), human papillomavirus (44, 96), respiratory syncytial virus (77), coxsackievirus B3 (19), Epstein-Barr virus (17, 50, 73), human cytomegalovirus (36, 37, 72), herpes simplex virus type 1 (7, 83), varicella-zoster virus (60), Kaposi''s sarcoma-associated herpesvirus (89), adenovirus (55), and simian virus 40 (SV40) (95). With this in mind, we also investigated whether the PI3K/Akt pathway played a pivotal role in orthopoxvirus biology. In this study, we show that the VACV- and CPXV-stimulated PI3K/Akt pathway not only contributes to the prevention of host-cell death but also plays a beneficial role in the viral replication cycle.  相似文献   

14.
The target of rapamycin (TOR) complex 1 (TORC1) signaling pathway is a critical regulator of translation and cell growth. To identify novel components of this pathway, we performed a kinome-wide RNA interference (RNAi) screen in Drosophila melanogaster S2 cells. RNAi targeting components of the p38 stress-activated kinase cascade prevented the cell size increase elicited by depletion of the TOR negative regulator TSC2. In mammalian and Drosophila tissue culture, as well as in Drosophila ovaries ex vivo, p38-activating stresses, such as H2O2 and anisomycin, were able to activate TORC1. This stress-induced TORC1 activation could be blocked by RNAi against mitogen-activated protein kinase kinase 3 and 6 (MKK3/6) or by the overexpression of dominant negative Rags. Interestingly, p38 was also required for the activation of TORC1 in response to amino acids and growth factors. Genetic ablation either of p38b or licorne, its upstream kinase, resulted in small flies consisting of small cells. Mutants with mutations in licorne or p38b are nutrition sensitive; low-nutrient food accentuates the small-organism phenotypes, as well as the partial lethality of the p38b null allele. These data suggest that p38 is an important positive regulator of TORC1 in both mammalian and Drosophila systems in response to certain stresses and growth factors.The target of rapamycin, TOR, is a highly conserved serine/threonine kinase that is a critical regulator of cell growth. It is a core component of two signaling complexes, TORC1 and TORC2 (60, 74). TORC1 is defined by the presence of Raptor in the complex, while TORC2 contains Rictor. Rictor and Raptor are mutually exclusive. Activation of the TORC1 pathway leads to increased protein translation, increased cell size, and increased proliferation, making this pathway an important target for emerging cancer therapies. Rapamycin is an inhibitor of TORC1 that is commonly used as an immunosuppressant following kidney transplantation (51). At least three analogs of rapamycin are currently being tested in solid and hematological tumors and have shown some promising results (21).The TORC1 pathway responds to numerous inputs, sensing both the desirability of and the capacity for growth. Many of these pathways control TORC1 signaling through phosphorylation of the tuberous sclerosis protein TSC2. TSC2 associates with TSC1 to form a heterodimeric GTPase-activating protein complex (GAP) that inactivates the small GTPase Rheb (24, 29, 67). While the exact molecular mechanism remains a topic of debate, activation of Rheb promotes the kinase activity of TORC1 (24, 29, 67). Rheb is required for the activation of TORC1 in response to both amino acids and growth factors (55, 62). In Drosophila melanogaster, mutation of either TOR or Rheb inhibits growth, leading to reduced body size and reduced cell size in mutant clones (42, 64). Mutation of either TSC1 or TSC2 has the predicted opposite effect, as tissue deficient for either of these proteins overgrows and contains large cells (49, 66).TORC1 is activated via the phosphatidylinositol 3′ kinase (PI3′K) pathway by growth-promoting mitogens, such as insulin and growth factors. Drosophila mutants with mutations of PI3′K pathway components have size phenotypes similar to those of the TOR and Rheb mutants (71). In mammalian cells, the PI3′K-mediated activation of TORC1 occurs at least in part through the phosphorylation of TSC2 by the PI3′K target AKT (30, 50). Interestingly, mutation of these residues in Drosophila has no impact on TSC2 function in vivo, suggesting that there may be other mechanisms through which PI3′K can activate Drosophila TOR (20). Recent work has suggested that the proline-rich AKT substrate PRAS40 may provide part of this link (23, 59, 69, 70). In addition, signaling through RAS activates extracellular signal-regulated kinase (ERK) and ribosomal S6 kinase (RSK), which can phosphorylate TSC2 and Raptor to activate TORC1 (13, 40, 56). There are also likely to be additional mechanisms through which growth factors activate Drosophila TOR that have not yet been identified.TORC1 activity is also controlled by the intracellular building blocks necessary to support cellular growth. The energy-sensing AMP-activated protein kinase (AMPK) pathway relays information about the energy status of the cell to TORC1 by phosphorylating TSC2. Unlike the inactivating phosphorylation of TSC2 by Akt, phosphorylation of TSC2 by AMPK promotes the GAP activity of the TSC complex (31). AMPK also phosphorylates Raptor, leading to decreased TORC1 activity (28). Thus, when energy levels are low, active AMPK inhibits TORC1.Amino acids also activate the TORC1 pathway, through a mechanism that requires Rheb, as well as the type III PI3′K VPS34 and the serine/threonine kinase mitogen-activated protein kinase kinase kinase kinase 3 (MAP4K3) (11, 22, 43). TORC1 thereby integrates information about the availability of amino acids and the amount of energy available for growth with growth factor signaling. Given its ancient function in adapting growth rates to environmental conditions, it is likely that TOR responds to a variety of stimuli, suggesting that many TOR control mechanisms remain to be uncovered. The Rag family of Ras-related small GTPases has recently been identified as a key component of the amino acid-sensing pathway, acting in parallel to Rheb (34, 58). Rag GTPases form heterodimers; RagA or RagB interacts with RagC or RagD. RagA and RagB are active when GTP bound, while RagC and RagD are active when bound to GDP (34, 58). Activation of the Rags by amino acids results in TOR relocalization to Rab7-containing vesicles (58). While the function of these vesicles in TORC1 signaling remains unclear, this relocalization is associated with increased TORC1 activity.TORC1 controls cell growth and translation through the phosphorylation and activation of components of the translational machinery, such as S6 kinase (S6K) and 4EBP1, an inhibitor of eukaryotic translation initiation factor 4E (eIF4E) activity (reviewed in reference 27). S6K phosphorylates the S6 ribosomal subunit, thereby increasing translation. Mice deficient for S6K1 are small and have small pancreatic beta cells and a correspondingly low level of circulating insulin (45). Mutation of the phosphorylation sites on S6 results in a similar phenotype, with small beta cells and fibroblasts (57). In Drosophila, mutation of S6K again reduces both cell and organism size (42), as does the overexpression of 4EBP (41).Interestingly, while mutation of the TORC1 pathway in mammalian cells reduces cell size by 10 to 15%, ablation of core TORC1 pathway components in Drosophila cells can affect cell size by up to 40% (73). In an attempt to identify novel components of the TORC1 pathway, we undertook an RNA interference (RNAi)-based screen of Drosophila S2 cells. We reasoned that the extreme size phenotypes observed in Drosophila cells upon TORC1 manipulations would facilitate the identification of modulators. In order to increase the likelihood of isolating novel regulators of TOR, we uncoupled TOR activity from many of its known nutritional controls by depleting TSC2 and screened for double-stranded RNAs (dsRNAs) that could reverse the cell size increase elicited by loss of TSC2. Depletion of multiple components of the p38 pathway was found to revert the TSC2 RNAi-induced cell size increase. Furthermore, activation of p38 is necessary and sufficient for the activation of TOR. Strikingly, mutation of components of the stress-activated p38 pathway in Drosophila has a similar phenotype to mutations in the TOR and insulin signaling pathway: a cell-autonomous cell size decrease, reduced body size, and a sensitization to the effects of nutritional stress.  相似文献   

15.
The serine/threonine kinase mammalian target of rapamycin (mTOR) governs growth, metabolism, and aging in response to insulin and amino acids (aa), and is often activated in metabolic disorders and cancer. Much is known about the regulatory signaling network that encompasses mTOR, but surprisingly few direct mTOR substrates have been established to date. To tackle this gap in our knowledge, we took advantage of a combined quantitative phosphoproteomic and interactomic strategy. We analyzed the insulin- and aa-responsive phosphoproteome upon inhibition of the mTOR complex 1 (mTORC1) component raptor, and investigated in parallel the interactome of endogenous mTOR. By overlaying these two datasets, we identified acinus L as a potential novel mTORC1 target. We confirmed acinus L as a direct mTORC1 substrate by co-immunoprecipitation and MS-enhanced kinase assays. Our study delineates a triple proteomics strategy of combined phosphoproteomics, interactomics, and MS-enhanced kinase assays for the de novo-identification of mTOR network components, and provides a rich source of potential novel mTOR interactors and targets for future investigation.The serine/threonine kinase mammalian target of rapamycin (mTOR)1 is conserved in all eukaryotes from yeast to mammals (1). mTOR is a central controller of cellular growth, whole body metabolism, and aging, and is frequently deregulated in metabolic diseases and cancer (2). Consequently, mTOR as well as its upstream and downstream cues are prime candidates for targeted drug development to alleviate the causes and symptoms of age-related diseases (3, 4). The identification of novel mTOR regulators and effectors thus remains a major goal in biomedical research. A vast body of literature describes a complex signaling network around mTOR. However, our current comparatively detailed knowledge of mTOR''s upstream cues contrasts with a rather limited set of known direct mTOR substrates.mTOR exists in two structurally and functionally distinct multiprotein complexes, termed mTORC1 and mTORC2. Both complexes contain mTOR kinase as well as the proteins mLST8 (mammalian lethal with SEC thirteen 8) (57), and deptor (DEP domain-containing mTOR-interacting protein) (8). mTORC1 contains the specific scaffold protein raptor (regulatory-associated protein of mTOR) (9, 10), whereas mTORC2 contains the specific binding partners rictor (rapamycin-insensitive companion of mTOR) (57), mSIN1 (TORC2 subunit MAPKAP1) (1113), and PRR5/L (proline rich protein 5/-like) (1416). The small macrolide rapamycin acutely inhibits mTORC1, but can also have long-term effects on mTORC2 (17, 18). More recently, ATP-analogs (19) that block both mTOR complexes, such as Torin 1 (20), have been developed. As rapamycin has already been available for several decades, our knowledge of signaling events associated with mTORC1 as well as its metabolic inputs and outputs is much broader as compared with mTORC2. mTORC1 responds to growth factors (insulin), nutrients (amino acids, aa) and energy (ATP). In response, mTORC1 activates anabolic processes (protein, lipid, nucleotide synthesis) and blocks catabolic processes (autophagy) to ultimately allow cellular growth (21). The insulin signal is transduced to mTORC1 via the insulin receptor (IR), and the insulin receptor substrate (IRS), which associates with class I phosphoinositide 3-kinases (PI3Ks). Subsequent phosphatidylinositol 3,4,5 trisphosphate (PIP3) binding leads to relocalization of the AGC kinases phosphoinositide-dependent protein kinase 1 (PDK1) and Akt (also termed protein kinase B, PKB) to the plasma membrane, where PDK1 phosphorylates Akt at T308 (22, 23). In response, Akt phosphorylates and inhibits the heterocomplex formed by the tuberous sclerosis complex proteins 1 and 2 (TSC1-TSC2) (24, 25). TSC1-TSC2 is the inhibitory, GTPase-activating protein for the mTORC1-inducing GTPase Ras homolog enriched in brain (rheb) (2630), which activates mTORC1 at the lysosome. mTORC1 localization depends on the presence of aa, which in a rag GTPase-dependent manner induce mTORC1 relocalization to lysosomes (31, 32). Low energy levels are sensed by the AMP-dependent kinase (AMPK), which in turn phosphorylates the TSC1-TSC2 complex (33) and raptor (34), thereby inhibiting mTORC1.mTORC1 phosphorylates its well-described downstream substrate S6-kinase (S6K) at T389, the proline-rich Akt substrate of 40 kDa (PRAS40) at S183, and the translational repressor 4E-binding protein (4E-BP) at T37/46 (3541). Unphosphorylated 4E-BP binds and inhibits the translation initiation factor 4G (eIF4G), which within the eIF4F complex mediates the scanning process of the ribosome to reach the start codon. Phosphorylation by mTORC1 inhibits 4E-BP''s interaction with eIF4E, thus allowing for assembly of eIF4F, and translation initiation (42, 43). More recently, also the IR-activating growth factor receptor-bound protein 10 (Grb10) (44, 45), the autophagy-initiating Unc-51-like kinase ULK1 (46), and the trifunctional enzymatic complex CAD composed of carbamoyl-phosphate synthetase 2, aspartate transcarbamoylase, and dihydroorotase (47, 48), which is required for nucleotide synthesis, have been described as direct mTORC1 substrates.mTORC2 activation is mostly described to be mediated by insulin, and this is mediated by a PI3K variant that is distinct from the PI3K upstream of mTORC1 (49, 50). Furthermore, mTORC2 responds to aa (5, 51). In response, mTORC2 phosphorylates the AGC kinases Akt at S473 (5255), and serum and glucocorticoid kinase SGK (56) and protein kinase C alpha (PKCalpha) (7) within their hydrophobic motifs (57, 58), to control cellular motility (57), hepatic glycolysis, and lipogenesis (59). In addition, mTOR autophosphorylation at S2481 has been established as an mTORC2 readout in several cell lines including HeLa cells (49).Given the multiplicity of effects via which mTOR controls cellular and organismal growth and metabolism, it is surprising that only relatively few direct mTOR substrates have been established to date. Proteomic studies are widely used to identify novel interactors and substrates of protein kinases. Two studies have recently shed light on the interaction of rapamycin and ATP-analog mTOR inhibitors with TSC2 inhibition in mammalian cells (44, 45), and one study has analyzed the effects of raptor and rictor knockouts in non-stimulated cells (48).In this work, we report a functional proteomics approach to study mTORC1 substrates. We used an inducible raptor knockdown to inhibit mTORC1 in HeLa cells, and analyzed the effect in combination with insulin and aa induction by quantitative phosphoproteomics using stable isotope labeling by amino acids in cell culture (SILAC) (60). In parallel, we purified endogenous mTOR complexes and studied the interactome of mTOR by SILAC-MS. Through comparative data evaluation, we identified acinus L as a potential novel aa/insulin-sensitive mTOR substrate. We further validated acinus L by co-immunoprecipitation and MS-enhanced kinase assays as a new direct mTORC1 substrate.  相似文献   

16.
17.
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   

18.
One essential downstream signaling pathway of receptor tyrosine kinases (RTKs), such as vascular endothelial growth factor receptor (VEGFR) and the Tie2 receptor, is the phosphoinositide-3 kinase (PI3K)-phosphoinositide-dependent protein kinase 1 (PDK1)-Akt/protein kinase B (PKB) cascade that plays a critical role in development and tumorigenesis. However, the role of PDK1 in cardiovascular development remains unknown. Here, we deleted PDK1 specifically in endothelial cells in mice. These mice displayed hemorrhage and hydropericardium and died at approximately embryonic day 11.5 (E11.5). Histological analysis revealed defective vascular remodeling and development and disrupted integrity between the endothelium and trabeculae/myocardium in the heart. The atrioventricular canal (AVC) cushion and valves failed to form, indicating a defect in epithelial-mesenchymal transition (EMT), together with increased endothelial apoptosis. Consistently, ex vivo AVC explant culture showed impeded mesenchymal outgrowth. Snail protein was reduced and was absent from the nucleus in AVC cells. Delivery of the Snail S6A mutant to the AVC explant effectively rescued EMT defects. Furthermore, adenoviral Akt delivery rescued EMT defects in AVC explant culture, and deletion of PTEN delayed embryonic lethality of PDK1 endothelial deletion mice by 1 day and rendered normal development of the AVC cushion in the PDK1-deficient heart. Taken together, these results have revealed an essential role of PDK1 in cardiovascular development through activation of Akt and Snail.Polypeptide growth factors, such as insulin, insulin-like growth factor 1 (IGF-I), vascular endothelial growth factor (VEGF), and angiopoietin 1 (Ang1), exert biological functions through binding to their transmembrane receptors that belong to a large family of receptor tyrosine kinases (RTKs) (4). Consequently, the receptor molecules form homo- or heterodimers, and the intracellular tyrosines at the carboxyl termini of the receptors become phosphorylated (37). Numerous distinct adaptor/regulatory proteins, through their Src homologous 2 (SH2) domains, bind to the phosphotyrosines and transduce the signal to downstream pathways, among which are two essential and well-characterized signaling cascades—the mitogen-activated protein kinase (MAPK) and phosphoinositide-3 kinase (PI3K)-phosphoinositide-dependent protein kinase 1 (PDK1)-Akt signaling pathways (4, 13, 37).The regulatory subunit of PI3K, p85, possesses the SH2 domain and can, therefore, bind to phosphotyrosines on the RTKs and subsequently render activation of the catalytic subunit of PI3K, p110 (7, 8). Active p110 phosphorylates phosphoinositide biphosphate (PIP2), turning it into PIP3 that recruits PDK1 and Akt to the cellular membrane, where Akt is phosphorylated at threonine 308 (T308 for Akt1) by PDK (5, 23, 30). The serine 473 (S473) of Akt (Akt1) is phosphorylated by mTOR complex 2 (mTORC2) and other kinases (17, 36). Phosphorylation of Akt at these two amino acids brings it to full activation. In PDK1-deficient embryonic stem (ES) cells, T308 phosphorylation was abolished and most of the Akt activity was lost, although the S473 phosphorylation was intact (40).Akt plays an important role in multiple biological processes, such as cell survival, growth, glucose metabolism, and angiogenesis (2, 12, 14-16, 22, 23, 39, 41-43). In mammals, there are three Akt isoforms, termed Akt 1, -2, and -3. Previously, we generated Akt1- and Akt3-deficient mice and studied their roles in mouse development (2, 15, 39, 42, 43). We found that the Akt1 and -3 double knockout (KO) (DKO) mice were embryonically lethal at around embryonic day 12 (E12) and manifested developmental defects in multiple tissues, including the cardiovascular system (14, 15, 43). These studies suggest that the Akt signaling pathway is involved in cardiovascular development.Other than Akt isoforms, PDK1 also activates another group of AGC family kinases, such as p70 ribosomal S6 kinase (S6K) (32), serum, and glucocorticoid-induced protein kinase (SGK) (26), p90 ribosomal S6 kinase (RSK) (21), and atypical isoforms of protein kinase C (PKC) (31). Comprehensive and intensive mouse genetic studies performed mainly by Alessi and coworkers have confirmed the regulation of these AGC kinases by PDK1 (3, 9, 10, 27-29, 40).PDK1 knockout mice were severely growth retarded and died at around E9.0, indicating an essential role of PDK1 in development (27). However, its function and downstream targets in cardiovascular development are still elusive. To study this, we deleted PDK1 specifically in endothelial cells through Cre recombinase-mediated excision (25). The results have revealed an essential role of PDK1 in vascular remodeling and integrity and in cardiac development through activation of Akt and its downstream target of Snail.  相似文献   

19.
20.
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