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 78 毫秒
1.
2.
3.
4.
5.
Mutations in SHP-2 phosphatase (PTPN11) that cause hyperactivation of its catalytic activity have been identified in Noonan syndrome and various childhood leukemias. Recent studies suggest that the gain-of-function (GOF) mutations of SHP-2 play a causal role in the pathogenesis of these diseases. However, the molecular mechanisms by which GOF mutations of SHP-2 induce these phenotypes are not fully understood. Here, we show that GOF mutations in SHP-2, such as E76K and D61G, drastically increase spreading and migration of various cell types, including hematopoietic cells, endothelial cells, and fibroblasts. More importantly, in vivo angiogenesis in SHP-2 D61G knock-in mice is also enhanced. Mechanistic studies suggest that the increased cell migration is attributed to the enhanced β1 integrin outside-in signaling. In response to β1 integrin cross-linking or fibronectin stimulation, activation of ERK and Akt kinases is greatly increased by SHP-2 GOF mutations. Also, integrin-induced activation of RhoA and Rac1 GTPases is elevated. Interestingly, mutant cells with the SHP-2 GOF mutation (D61G) are more sensitive than wild-type cells to the suppression of cell motility by inhibition of these pathways. Collectively, these studies reaffirm the positive role of SHP-2 phosphatase in cell motility and suggest a new mechanism by which SHP-2 GOF mutations contribute to diseases.SHP-2, a multifunctional SH2 domain-containing protein-tyrosine phosphatase implicated in diverse cell signaling processes (13), plays a critical role in cellular function. Homozygous deletion of Exon 2 (4) or Exon 3 (5) of the SHP-2 gene (PTPN11) in mice leads to early embryonic lethality prior to and at midgestation, respectively. SHP-2 null mutant mice die much earlier, at peri-implantation (4). Exon 3 deletion mutation of SHP-2 blocks hematopoietic potential of embryonic stem cells both in vitro and in vivo (68), whereas SHP-2 null mutation causes inner cell mass death and diminished trophoblast stem cell survival (4). Recent studies on SHP-2 conditional knock-out or tissue-specific knock-out mice have further revealed an array of important functions of this phosphatase in various physiological processes (912). The phenotypes demonstrated by loss of SHP-2 function are apparently attributed to the role of SHP-2 in the cell signaling pathways induced by growth factors/cytokines. SHP-2 generally promotes signal transmission in growth factor/cytokine signaling in both catalytic-dependent and -independent fashion (13). The positive role of SHP-2 in the intracellular signaling processes, in particular, the ERK3 and PI3K/Akt kinase pathways, has been well established, although the underlying mechanism remains elusive, in particular, the signaling function of the catalytic activity of SHP-2 in these pathways is poorly understood.In addition to the role of SHP-2 in cell proliferation and differentiation, the phenotypes induced by loss of SHP-2 function may be associated with its role in cell migration. Indeed, dominant negative SHP-2 disrupts Xenopus gastrulation, causing tail truncations (13, 14). Targeted Exon 3 deletion mutation in SHP-2 results in decreased cell spreading, migration (15, 16), and impaired limb development in the chimeric mice (7). The role of SHP-2 in cell adhesion and migration has also been demonstrated by catalytically inactive mutant SHP-2-overexpressing cells (1720). The molecular mechanisms by which SHP-2 regulates these cellular processes, however, have not been well defined. For example, the role of SHP-2 in the activation of the Rho family small GTPases that is critical for cell motility is still controversial. Both positive (19, 21, 22) and negative roles (18, 23) for SHP-2 in this context have been reported. Part of the reason for this discrepancy might be due to the difference in the cell models used. Catalytically inactive mutant SHP-2 was often used to determine the role of SHP-2 in cell signaling. In the catalytically inactive mutant SHP-2-overexpressing cells, the catalytic activity of endogenous SHP-2 is inhibited. However, as SHP-2 also functions independent of its catalytic activity, overexpression of catalytically deficient SHP-2 may also increase its scaffolding function, generating complex effects.The critical role of SHP-2 in cellular function is further underscored by the identification of SHP-2 mutations in human diseases. Genetic lesions in PTPN11 that cause hyperactivation of SHP-2 catalytic activity have been identified in the developmental disorder Noonan syndrome (24) and various childhood leukemias, including juvenile myelomonocytic leukemia (JMML), B cell acute lymphoblastic leukemia, and acute myeloid leukemia (25, 26). In addition, activating mutations in SHP-2 have been identified in sporadic solid tumors (27). The SHP-2 mutations appear to play a causal role in the development of these diseases as SHP-2 mutations and other JMML-associated Ras or Neurofibromatosis 1 mutations are mutually exclusive in the patients (2427). Moreover, single SHP-2 gain-of-function (GOF) mutations are sufficient to induce Noonan syndrome, cytokine hypersensitivity in hematopoietic progenitor cells, and JMML-like myeloproliferative disease in mice (2832). Gain-of-function cell models derived from the newly available SHP-2 GOF mutation (D61G) knock-in mice (28) now provide us with a good opportunity to clarify the role of SHP-2 in cell motility. Unlike the dominant negative approach in which overexpression of mutant forms of SHP-2 generates complex effects, the SHP-2 D61G knock-in model eliminates this possibility as the mutant SHP-2 is expressed at the physiological level (28). Additionally, defining signaling functions of GOF mutant SHP-2 in cell movement can also help elucidate the molecular mechanisms by which SHP-2 mutations contribute to the relevant diseases.  相似文献   

6.
Wnt11 signals through both canonical (β-catenin) and non-canonical pathways and is up-regulated during osteoblast differentiation and fracture healing. In these studies, we evaluated the role of Wnt11 during osteoblastogenesis. Wnt11 overexpression in MC3T3E1 pre-osteoblasts increases β-catenin accumulation and promotes bone morphogenetic protein (BMP)-induced expression of alkaline phosphatase and mineralization. Wnt11 dramatically increases expression of the osteoblast-associated genes Dmp1 (dentin matrix protein 1), Phex (phosphate-regulating endopeptidase homolog), and Bsp (bone sialoprotein). Wnt11 also increases expression of Rspo2 (R-spondin 2), a secreted factor known to enhance Wnt signaling. Overexpression of Rspo2 is sufficient for increasing Dmp1, Phex, and Bsp expression and promotes bone morphogenetic protein-induced mineralization. Knockdown of Rspo2 abrogates Wnt11-mediated osteoblast maturation. Antagonism of T-cell factor (Tcf)/β-catenin signaling with dominant negative Tcf blocks Wnt11-mediated expression of Dmp1, Phex, and Rspo2 and decreases mineralization. However, dominant negative Tcf fails to block the osteogenic effects of Rspo2 overexpression. These studies show that Wnt11 signals through β-catenin, activating Rspo2 expression, which is then required for Wnt11-mediated osteoblast maturation.Wnt signaling is a key regulator of osteoblast differentiation and maturation. In mesenchymal stem cell lines, canonical Wnt signaling by Wnt10b enhances osteoblast differentiation (1). Canonical Wnt signaling through β-catenin has also been shown to enhance the chondroinductive and osteoinductive properties of BMP22 (2, 3). During BMP2-induced osteoblast differentiation of mesenchymal stem cell lines, cross-talk between BMP and Wnt pathways converges through the interaction of Smad4 with β-catenin (2).Canonical Wnt signaling is also critical for skeletal development and homeostasis. During limb development, expression of Wnt3a in the apical ectodermal ridge of limb buds maintains cells in a highly proliferative and undifferentiated state (4, 5). Disruption of canonical Wnt signaling in Lrp5/Lrp6 compound knock-out mice results in limb- and digit-patterning defects (6). Wnt signaling is also involved in the maintenance of post-natal bone mass. Gain of function in the Wnt co-receptor Lrp5 leads to increased bone mass, whereas loss of Lrp5 function is associated with decreased bone mass and osteoporosis pseudoglioma syndrome (7, 8). Mice with increased Wnt10b expression have increased trabecular bone, whereas Wnt10b-deficient mice have reduced trabecular bone (9). Similarly, mice nullizygous for the Wnt antagonist sFrp1 have increased trabecular bone accrual throughout adulthood (10).Although canonical Wnt signaling regulates osteoblastogenesis and bone formation, the profile of endogenous Wnts that play a role in osteoblast differentiation and maturation is not well described. During development, Wnt11 is expressed in the perichondrium and in the axial skeleton and sternum (11). Wnt11 expression is increased during glucocorticoid-induced osteogenesis (12), indicating a potential role for Wnt11 in osteoblast differentiation. Interestingly, Wnt11 activates both β-catenin-dependent as well as β-catenin-independent signaling pathways (13). Targeted disruption of Wnt11 results in late embryonic/early post-natal death because of cardiac dysfunction (14). Although these mice have no reported skeletal developmental abnormalities, early lethality obfuscates a detailed examination of post-natal skeletal modeling and remodeling.In murine development, Wnt11 expression overlaps with the expression of R-spondin 2 (Rspo2) in the apical ectodermal ridge (11, 15). R-spondins are a novel family of proteins that share structural features, including two conserved cysteinerich furin-like domains and a thrombospondin type I repeat (16). The four R-spondin family members can activate canonical Wnt signaling (15, 1719). Rspo3 interacts with Frizzled 8 and Lrp6 and enhances Wnt ligand signaling. Rspo1 enhances Wnt signaling by interacting with Lrp6 and inhibiting Dkk-mediated receptor internalization (20). Rspo1 was also shown to potentiate Wnt3a-mediated osteoblast differentiation (21). Rspo2 knock-out mice, which die at birth, have limb patterning defects associated with altered β-catenin signaling (2224). However, the role of Rspo2 in osteoblast differentiation and maturation remains unclear.Herein we report that Wnt11 overexpression in MC3T3E1 pre-osteoblasts activates β-catenin and augments BMP-induced osteoblast maturation and mineralization. Wnt11 increases the expression of Rspo2. Overexpression of Rspo2 in MC3T3E1 is sufficient for augmenting BMP-induced osteoblast maturation and mineralization. Although antagonism of Tcf/β-catenin signaling blocks the osteogenic effects of Wnt11, Rspo2 rescues this block, and knockdown of Rspo2 shows that it is required for Wnt11-mediated osteoblast maturation and mineralization. These studies identify both Wnt11 and Rspo2 as novel mediators of osteoblast maturation and mineralization.  相似文献   

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

8.
NHE5 is a brain-enriched Na+/H+ exchanger that dynamically shuttles between the plasma membrane and recycling endosomes, serving as a mechanism that acutely controls the local pH environment. In the current study we show that secretory carrier membrane proteins (SCAMPs), a group of tetraspanning integral membrane proteins that reside in multiple secretory and endocytic organelles, bind to NHE5 and co-localize predominantly in the recycling endosomes. In vitro protein-protein interaction assays revealed that NHE5 directly binds to the N- and C-terminal cytosolic extensions of SCAMP2. Heterologous expression of SCAMP2 but not SCAMP5 increased cell-surface abundance as well as transporter activity of NHE5 across the plasma membrane. Expression of a deletion mutant lacking the SCAMP2-specific N-terminal cytosolic domain, and a mini-gene encoding the N-terminal extension, reduced the transporter activity. Although both Arf6 and Rab11 positively regulate NHE5 cell-surface targeting and NHE5 activity across the plasma membrane, SCAMP2-mediated surface targeting of NHE5 was reversed by dominant-negative Arf6 but not by dominant-negative Rab11. Together, these results suggest that SCAMP2 regulates NHE5 transit through recycling endosomes and promotes its surface targeting in an Arf6-dependent manner.Neurons and glial cells in the central and peripheral nervous systems are especially sensitive to perturbations of pH (1). Many voltage- and ligand-gated ion channels that control membrane excitability are sensitive to changes in cellular pH (1-3). Neurotransmitter release and uptake are also influenced by cellular and organellar pH (4, 5). Moreover, the intra- and extracellular pH of both neurons and glia are modulated in a highly transient and localized manner by neuronal activity (6, 7). Thus, neurons and glia require sophisticated mechanisms to finely tune ion and pH homeostasis to maintain their normal functions.Na+/H+ exchangers (NHEs)3 were originally identified as a class of plasma membrane-bound ion transporters that exchange extracellular Na+ for intracellular H+, and thereby regulate cellular pH and volume. Since the discovery of NHE1 as the first mammalian NHE (8), eight additional isoforms (NHE2-9) that share 25-70% amino acid identity have been isolated in mammals (9, 10). NHE1-5 commonly exhibit transporter activity across the plasma membrane, whereas NHE6-9 are mostly found in organelle membranes and are believed to regulate organellar pH in most cell types at steady state (11). More recently, NHE10 was identified in human and mouse osteoclasts (12, 13). However, the cDNA encoding NHE10 shares only a low degree of sequence similarity with other known members of the NHE gene family, raising the possibility that this sodium-proton exchanger may belong to a separate gene family distantly related to NHE1-9 (see Ref. 9).NHE gene family members contain 12 putative transmembrane domains at the N terminus followed by a C-terminal cytosolic extension that plays a role in regulation of the transporter activity by protein-protein interactions and phosphorylation. NHEs have been shown to regulate the pH environment of synaptic nerve terminals and to regulate the release of neurotransmitters from multiple neuronal populations (14-16). The importance of NHEs in brain function is further exemplified by the findings that spontaneous or directed mutations of the ubiquitously expressed NHE1 gene lead to the progression of epileptic seizures, ataxia, and increased mortality in mice (17, 18). The progression of the disease phenotype is associated with loss of specific neuron populations and increased neuronal excitability. However, NHE1-null mice appear to develop normally until 2 weeks after birth when symptoms begin to appear. Therefore, other mechanisms may compensate for the loss of NHE1 during early development and play a protective role in the surviving neurons after the onset of the disease phenotype.NHE5 was identified as a unique member of the NHE gene family whose mRNA is expressed almost exclusively in the brain (19, 20), although more recent studies have suggested that NHE5 might be functional in other cell types such as sperm (21, 22) and osteosarcoma cells (23). Curiously, mutations found in several forms of congenital neurological disorders such as spinocerebellar ataxia type 4 (24-26) and autosomal dominant cerebellar ataxia (27-29) have been mapped to chromosome 16q22.1, a region containing NHE5. However, much remains unknown as to the molecular regulation of NHE5 and its role in brain function.Very few if any proteins work in isolation. Therefore identification and characterization of binding proteins often reveal novel functions and regulation mechanisms of the protein of interest. To begin to elucidate the biological role of NHE5, we have started to explore NHE5-binding proteins. Previously, β-arrestins, multifunctional scaffold proteins that play a key role in desensitization of G-protein-coupled receptors, were shown to directly bind to NHE5 and promote its endocytosis (30). This study demonstrated that NHE5 trafficking between endosomes and the plasma membrane is regulated by protein-protein interactions with scaffold proteins. More recently, we demonstrated that receptor for activated C-kinase 1 (RACK1), a scaffold protein that links signaling molecules such as activated protein kinase C, integrins, and Src kinase (31), directly interacts with and activates NHE5 via integrin-dependent and independent pathways (32). These results further indicate that NHE5 is partly associated with focal adhesions and that its targeting to the specialized microdomain of the plasma membrane may be regulated by various signaling pathways.Secretory carrier membrane proteins (SCAMPs) are a family of evolutionarily conserved tetra-spanning integral membrane proteins. SCAMPs are found in multiple organelles such as the Golgi apparatus, trans-Golgi network, recycling endosomes, synaptic vesicles, and the plasma membrane (33, 34) and have been shown to play a role in exocytosis (35-38) and endocytosis (39). Currently, five isoforms of SCAMP have been identified in mammals. The extended N terminus of SCAMP1-3 contain multiple Asn-Pro-Phe (NPF) repeats, which may allow these isoforms to participate in clathrin coat assembly and vesicle budding by binding to Eps15 homology (EH)-domain proteins (40, 41). Further, SCAMP2 was shown recently to bind to the small GTPase Arf6 (38), which is believed to participate in traffic between the recycling endosomes and the cell surface (42, 43). More recent studies have suggested that SCAMPs bind to organellar membrane type NHE7 (44) and the serotonin transporter SERT (45) and facilitate targeting of these integral membrane proteins to specific intracellular compartments. We show in the current study that SCAMP2 binds to NHE5, facilitates the cell-surface targeting of NHE5, and elevates Na+/H+ exchange activity at the plasma membrane, whereas expression of a SCAMP2 deletion mutant lacking the N-terminal domain containing the NPF repeats suppresses the effect. Further we show that this activity of SCAMP2 requires an active form of a small GTPase Arf6, but not Rab11. We propose a model in which SCAMPs bind to NHE5 in the endosomal compartment and control its cell-surface abundance via an Arf6-dependent pathway.  相似文献   

9.
10.
11.
12.
13.
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   

14.
Thimet oligopeptidase (EC 3.4.24.15; EP24.15) is an intracellular enzyme that has been proposed to metabolize peptides within cells, thereby affecting antigen presentation and G protein-coupled receptor signal transduction. However, only a small number of intracellular substrates of EP24.15 have been reported previously. Here we have identified over 100 peptides in human embryonic kidney 293 (HEK293) cells that are derived from intracellular proteins; many but not all of these peptides are substrates or products of EP24.15. First, cellular peptides were extracted from HEK293 cells and incubated in vitro with purified EP24.15. Then the peptides were labeled with isotopic tags and analyzed by mass spectrometry to obtain quantitative data on the extent of cleavage. A related series of experiments tested the effect of overexpression of EP24.15 on the cellular levels of peptides in HEK293 cells. Finally, synthetic peptides that corresponded to 10 of the cellular peptides were incubated with purified EP24.15 in vitro, and the cleavage was monitored by high pressure liquid chromatography and mass spectrometry. Many of the EP24.15 substrates identified by these approaches are 9–11 amino acids in length, supporting the proposal that EP24.15 can function in the degradation of peptides that could be used for antigen presentation. However, EP24.15 also converts some peptides into products that are 8–10 amino acids, thus contributing to the formation of peptides for antigen presentation. In addition, the intracellular peptides described here are potential candidates to regulate protein interactions within cells.Intracellular protein turnover is a crucial step for cell functioning, and if this process is impaired, the elevated levels of aged proteins usually lead to the formation of intracellular insoluble aggregates that can cause severe pathologies (1). In mammalian cells, most proteins destined for degradation are initially tagged with a polyubiquitin chain in an energy-dependent process and then digested to small peptides by the 26 S proteasome, a large proteolytic complex involved in the regulation of cell division, gene expression, and other key processes (2, 3). In eukaryotes, 30–90% of newly synthesized proteins may be degraded by proteasomes within minutes of synthesis (3, 4). In addition to proteasomes, other extralysosomal proteolytic systems have been reported (5, 6). The proteasome cleaves proteins into peptides that are typically 2–20 amino acids in length (7). In most cases, these peptides are thought to be rapidly hydrolyzed into amino acids by aminopeptidases (810). However, some intracellular peptides escape complete degradation and are imported into the endoplasmic reticulum where they associate with major histocompatibility complex class I (MHC-I)3 molecules and traffic to the cell surface for presentation to the immune system (1012). Additionally, based on the fact that free peptides added to the intracellular milieu can regulate cellular functions mediated by protein interactions such as gene regulation, metabolism, cell signaling, and protein targeting (13, 14), intracellular peptides generated by proteasomes that escape degradation have been suggested to play a role in regulating protein interactions (15). Indeed, oligopeptides isolated from rat brain tissue using the catalytically inactive EP24.15 (EC 3.4.24.15) were introduced into Chinese hamster ovarian-S and HEK293 cells and were found capable of altering G protein-coupled receptor signal transduction (16). Moreover, EP24.15 overexpression itself changed both angiotensin II and isoproterenol signal transduction, suggesting a physiological function for its intracellular substrates/products (16).EP24.15 is a zinc-dependent peptidase of the metallopeptidase M3 family that contains the HEXXH motif (17). This enzyme was first described as a neuropeptide-degrading enzyme present in the soluble fraction of brain homogenates (18). Whereas EP24.15 can be secreted (19, 20), its predominant location in the cytosol and nucleus suggests that the primary function of this enzyme is not the extracellular degradation of neuropeptides and hormones (21, 22). EP24.15 was shown in vivo to participate in antigen presentation through MHC-I (2325) and in vitro to bind (26) or degrade (27) some MHC-I associated peptides. EP24.15 has also been shown in vitro to degrade peptides containing 5–17 amino acids produced after proteasome digestion of β-casein (28). EP24.15 shows substrate size restriction to peptides containing from 5 to 17 amino acids because of its catalytic center that is located in a deep channel (29). Despite the size restriction, EP24.15 has a broad substrate specificity (30), probably because a significant portion of the enzyme-binding site is lined with potentially flexible loops that allow reorganization of the active site following substrate binding (29). Recently, it has also been suggested that certain substrates may be cleaved by an open form of EP24.15 (31). This characteristic is supported by the ability of EP24.15 to accommodate different amino acid residues at subsites S4 to S3′, which even includes the uncommon post-proline cleavage (30). Such biochemical and structural features make EP24.15 a versatile enzyme to degrade structurally unrelated oligopeptides.Previously, brain peptides that bound to catalytically inactive EP24.15 were isolated and identified using mass spectrometry (22). The majority of peptides captured by the inactive enzyme were intracellular protein fragments that efficiently interacted with EP24.15; the smallest peptide isolated in these assays contained 5 and the largest 17 amino acids (15, 16, 22, 32), which is within the size range previously reported for natural and synthetic substrates of EP24.15 (18, 30, 33, 34). Interestingly, the peptides released by the proteasome are in the same size range of EP24.15 competitive inhibitors/substrates (7, 35, 36). Taken altogether, these data suggest that in the intracellular environment EP24.15 could further cleave proteasome-generated peptides unrelated to MHC-I antigen presentation (15).Although the mutated inactive enzyme “capture” assay was successful in identifying several cellular protein fragments that were substrates for EP24.15, it also found some interacting peptides that were not substrates. In this study, we used several approaches to directly screen for cellular peptides that were cleaved by EP24.15. The first approach involved the extraction of cellular peptides from the HEK293 cell line, incubation in vitro with purified EP24.15, labeling with isotopic tags, and analysis by mass spectrometry to obtain quantitative data on the extent of cleavage. The second approach examined the effect of EP24.15 overexpression on the cellular levels of peptides in the HEK293 cell line. The third set of experiments tested synthetic peptides with purified EP24.15 in vitro, and examined cleavage by high pressure liquid chromatography and mass spectrometry. Collectively, these studies have identified a large number of intracellular peptides, including those that likely represent the endogenous substrates and products of EP24.15, and this original information contributes to a better understanding of the function of this enzyme in vivo.  相似文献   

15.
Formin-homology (FH) 2 domains from formin proteins associate processively with the barbed ends of actin filaments through many rounds of actin subunit addition before dissociating completely. Interaction of the actin monomer-binding protein profilin with the FH1 domain speeds processive barbed end elongation by FH2 domains. In this study, we examined the energetic requirements for fast processive elongation. In contrast to previous proposals, direct microscopic observations of single molecules of the formin Bni1p from Saccharomyces cerevisiae labeled with quantum dots showed that profilin is not required for formin-mediated processive elongation of growing barbed ends. ATP-actin subunits polymerized by Bni1p and profilin release the γ-phosphate of ATP on average >2.5 min after becoming incorporated into filaments. Therefore, the release of γ-phosphate from actin does not drive processive elongation. We compared experimentally observed rates of processive elongation by a number of different FH2 domains to kinetic computer simulations and found that actin subunit addition alone likely provides the energy for fast processive elongation of filaments mediated by FH1FH2-formin and profilin. We also studied the role of FH2 structure in processive elongation. We found that the flexible linker joining the two halves of the FH2 dimer has a strong influence on dissociation of formins from barbed ends but only a weak effect on elongation rates. Because formins are most vulnerable to dissociation during translocation along the growing barbed end, we propose that the flexible linker influences the lifetime of this translocative state.Formins are multidomain proteins that assemble unbranched actin filament structures for diverse processes in eukaryotic cells (reviewed in Ref. 1). Formins stimulate nucleation of actin filaments and, in the presence of the actin monomer-binding protein profilin, speed elongation of the barbed ends of filaments (2-6). The ability of formins to influence elongation depends on the ability of single formin molecules to remain bound to a growing barbed end through multiple rounds of actin subunit addition (7, 8). To stay associated during subunit addition, a formin molecule must translocate processively on the barbed end as each actin subunit is added (1, 9-12). This processive elongation of a barbed end by a formin is terminated when the formin dissociates stochastically from the growing end during translocation (4, 10).The formin-homology (FH)2 1 and 2 domains are the best conserved domains of formin proteins (2, 13, 14). The FH2 domain is the signature domain of formins, and in many cases, is sufficient for both nucleation and processive elongation of barbed ends (2-4, 7, 15). Head-to-tail homodimers of FH2 domains (12, 16) encircle the barbed ends of actin filaments (9). In vitro, association of barbed ends with FH2 domains slows elongation by limiting addition of free actin monomers. This “gating” behavior is usually explained by a rapid equilibrium of the FH2-associated end between an open state competent for actin monomer association and a closed state that blocks monomer binding (4, 9, 17).Proline-rich FH1 domains located N-terminal to FH2 domains are required for profilin to stimulate formin-mediated elongation. Individual tracks of polyproline in FH1 domains bind 1:1 complexes of profilin-actin and transfer the actin directly to the FH2-associated barbed end to increase processive elongation rates (4-6, 8, 10, 17).Rates of elongation and dissociation from growing barbed ends differ widely for FH1FH2 fragments from different formin homologs (4). We understand few aspects of FH1FH2 domains that influence gating, elongation or dissociation. In this study, we examined the source of energy for formin-mediated processive elongation, and the influence of FH2 structure on elongation and dissociation from growing ends. In contrast to previous proposals (6, 18), we found that fast processive elongation mediated by FH1FH2-formins is not driven by energy from the release of the γ-phosphate from ATP-actin filaments. Instead, the data show that the binding of an actin subunit to the barbed end provides the energy for processive elongation. We found that in similar polymerizing conditions, different natural FH2 domains dissociate from growing barbed ends at substantially different rates. We further observed that the length of the flexible linker between the subunits of a FH2 dimer influences dissociation much more than elongation.  相似文献   

16.
Aggregation of the Ure2 protein is at the origin of the [URE3] prion trait in the yeast Saccharomyces cerevisiae. The N-terminal region of Ure2p is necessary and sufficient to induce the [URE3] phenotype in vivo and to polymerize into amyloid-like fibrils in vitro. However, as the N-terminal region is poorly ordered in the native state, making it difficult to detect structural changes in this region by spectroscopic methods, detailed information about the fibril assembly process is therefore lacking. Short fibril-forming peptide regions (4–7 residues) have been identified in a number of prion and other amyloid-related proteins, but such short regions have not yet been identified in Ure2p. In this study, we identify a unique cysteine mutant (R17C) that can greatly accelerate the fibril assembly kinetics of Ure2p under oxidizing conditions. We found that the segment QVNI, corresponding to residues 18–21 in Ure2p, plays a critical role in the fast assembly properties of R17C, suggesting that this segment represents a potential amyloid-forming region. A series of peptides containing the QVNI segment were found to form fibrils in vitro. Furthermore, the peptide fibrils could seed fibril formation for wild-type Ure2p. Preceding the QVNI segment with a cysteine or a hydrophobic residue, instead of a charged residue, caused the rate of assembly into fibrils to increase greatly for both peptides and full-length Ure2p. Our results indicate that the potential amyloid stretch and its preceding residue can modulate the fibril assembly of Ure2p to control the initiation of prion formation.The [URE3] phenotype of Saccharomyces cerevisiae arises because of conversion of the Ure2 protein to an aggregated propagatable prion state (1, 2). Ure2p contains two regions: a poorly structured N-terminal region and a compactly folded C-terminal region (3, 4). The N-terminal region is rich in Asn and Gln residues, is highly flexible, and is without any detectable ordered secondary structure (46). This region is necessary and sufficient for prion behavior in vivo (2) and amyloid-forming capacity in vitro (5, 7), so it is referred to as the prion domain (PrD).2 The C-terminal region has a fold similar to the glutathione S-transferase superfamily (8, 9) and possesses glutathione-dependent peroxidase activity (10). Upon fibril formation, the N-terminal region undergoes a significant conformational change from an unfolded to a thermally resistant conformation (11), whereas the glutathione S-transferase-like C-terminal domain retains its enzymatic activity, suggesting that little conformational change occurs (10, 12). Ure2p fibrils show various morphologies, including variations in thickness and the presence or absence of a periodic twist (1316). The overall structure of the fibrils imaged by cryoelectron microscopy suggests that the intact fibrils contain a 4-nm amyloid filament backbone surrounded by C-terminal globular domains (17).It is widely accepted that disulfide bonds play a critical role in maintaining protein stability (1821) and also affect the process of protein folding by influencing the folding pathway (2225). A recent study shows that the presence of a disulfide bond in a protein can markedly accelerate the folding process (26). Therefore, a disulfide bond is a useful tool to study protein folding. In the study of prion and other amyloid-related proteins, cysteine scanning has been widely used to study the structure of amyloid fibrils, the driving force of amyloid formation, and the plasticity of amyloid fibrils (13, 2731).Short segments from amyloid-related proteins, including IAPP (islet amyloid polypeptide), β2-microglobulin, insulin, and the amyloid-β peptide, show amyloid-forming capacity (3234). Hence, the amyloid stretch hypothesis has been proposed, which suggests that a short amino acid stretch bearing a highly amyloidogenic motif might supply most of the driving force needed to trigger the self-catalytic assembly process of a protein to form fibrils (35, 36). In support of this hypothesis, it was found that the insertion of an amyloidogenic stretch into a non-amyloid-related protein can trigger the amyloidosis of the protein (36). At the same time, the structural information obtained from microcrystals formed by amyloidogenic stretches and bearing cross-β-structure has contributed significantly to our understanding of the structure of intact fibrils at the atomic level (34, 37). However, no amyloidogenic stretches <10 amino acids have so far been identified in the yeast prion protein Ure2.In this study, we performed a cysteine scan within the N-terminal PrD of Ure2p and found a unique cysteine mutant (R17C) that eliminates the lag phase of the Ure2p fibril assembly reaction upon the addition of oxidizing agents. Furthermore, we identified a 4-residue region adjacent to Arg17 as a potential amyloid stretch in Ure2p.  相似文献   

17.
Intersectin-short (intersectin-s) is a multimodule scaffolding protein functioning in constitutive and regulated forms of endocytosis in non-neuronal cells and in synaptic vesicle (SV) recycling at the neuromuscular junction of Drosophila and Caenorhabditis elegans. In vertebrates, alternative splicing generates a second isoform, intersectin-long (intersectin-l), that contains additional modular domains providing a guanine nucleotide exchange factor activity for Cdc42. In mammals, intersectin-s is expressed in multiple tissues and cells, including glia, but excluded from neurons, whereas intersectin-l is a neuron-specific isoform. Thus, intersectin-I may regulate multiple forms of endocytosis in mammalian neurons, including SV endocytosis. We now report, however, that intersectin-l is localized to somatodendritic regions of cultured hippocampal neurons, with some juxtanuclear accumulation, but is excluded from synaptophysin-labeled axon terminals. Consistently, intersectin-l knockdown (KD) does not affect SV recycling. Instead intersectin-l co-localizes with clathrin heavy chain and adaptor protein 2 in the somatodendritic region of neurons, and its KD reduces the rate of transferrin endocytosis. The protein also co-localizes with F-actin at dendritic spines, and intersectin-l KD disrupts spine maturation during development. Our data indicate that intersectin-l is indeed an important regulator of constitutive endocytosis and neuronal development but that it is not a prominent player in the regulated endocytosis of SVs.Clathrin-mediated endocytosis (CME)4 is a major mechanism by which cells take up nutrients, control the surface levels of multiple proteins, including ion channels and transporters, and regulate the coupling of signaling receptors to downstream signaling cascades (1-5). In neurons, CME takes on additional specialized roles; it is an important process regulating synaptic vesicle (SV) availability through endocytosis and recycling of SV membranes (6, 7), it shapes synaptic plasticity (8-10), and it is crucial in maintaining synaptic membranes and membrane structure (11).Numerous endocytic accessory proteins participate in CME, interacting with each other and with core components of the endocytic machinery such as clathrin heavy chain (CHC) and adaptor protein-2 (AP-2) through specific modules and peptide motifs (12). One such module is the Eps15 homology domain that binds to proteins bearing NPF motifs (13, 14). Another is the Src homology 3 (SH3) domain, which binds to proline-rich domains in protein partners (15). Intersectin is a multimodule scaffolding protein that interacts with a wide range of proteins, including several involved in CME (16). Intersectin has two N-terminal Eps15 homology domains that are responsible for binding to epsin, SCAMP1, and numb (17-19), a central coil-coiled domain that interacts with Eps15 and SNAP-23 and -25 (17, 20, 21), and five SH3 domains in its C-terminal region that interact with multiple proline-rich domain proteins, including synaptojanin, dynamin, N-WASP, CdGAP, and mSOS (16, 22-25). The rich binding capability of intersectin has linked it to various functions from CME (17, 26, 27) and signaling (22, 28, 29) to mitogenesis (30, 31) and regulation of the actin cytoskeleton (23).Intersectin functions in SV recycling at the neuromuscular junction of Drosophila and C. elegans where it acts as a scaffold, regulating the synaptic levels of endocytic accessory proteins (21, 32-34). In vertebrates, the intersectin gene is subject to alternative splicing, and a longer isoform (intersectin-l) is generated that is expressed exclusively in neurons (26, 28, 35, 36). This isoform has all the binding modules of its short (intersectin-s) counterpart but also has additional domains: a DH and a PH domain that provide guanine nucleotide exchange factor (GEF) activity specific for Cdc42 (23, 37) and a C2 domain at the C terminus. Through its GEF activity and binding to actin regulatory proteins, including N-WASP, intersectin-l has been implicated in actin regulation and the development of dendritic spines (19, 23, 24). In addition, because the rest of the binding modules are shared between intersectin-s and -l, it is generally thought that the two intersectin isoforms have the same endocytic functions. In particular, given the well defined role for the invertebrate orthologs of intersectin-s in SV endocytosis, it is thought that intersectin-l performs this role in mammalian neurons, which lack intersectin-s. Defining the complement of intersectin functional activities in mammalian neurons is particularly relevant given that the protein is involved in the pathophysiology of Down syndrome (DS). Specifically, the intersectin gene is localized on chromosome 21q22.2 and is overexpressed in DS brains (38). Interestingly, alterations in endosomal pathways are a hallmark of DS neurons and neurons from the partial trisomy 16 mouse, Ts65Dn, a model for DS (39, 40). Thus, an endocytic trafficking defect may contribute to the DS disease process.Here, the functional roles of intersectin-l were studied in cultured hippocampal neurons. We find that intersectin-l is localized to the somatodendritic regions of neurons, where it co-localizes with CHC and AP-2 and regulates the uptake of transferrin. Intersectin-l also co-localizes with actin at dendritic spines and disrupting intersectin-l function alters dendritic spine development. In contrast, intersectin-l is absent from presynaptic terminals and has little or no role in SV recycling.  相似文献   

18.
p53 is an important tumor suppressor regulating the cell cycle at multiple stages in higher vertebrates. The p53 gene is frequently deleted or mutated in human cancers, resulting in loss of p53 activity. This leads to centrosome amplification, aneuploidy, and tumorigenesis, three phenotypes also observed after overexpression of the oncogenic kinase Aurora A. Accordingly, recent studies have focused on the relationship between these two proteins. p53 and Aurora A have been reported to interact in mammalian cells, but the function of this interaction remains unclear. We recently reported that Xenopus p53 can inhibit Aurora A activity in vitro but only in the absence of TPX2. Here we investigate the interplay between Xenopus Aurora A, TPX2, and p53 and show that newly synthesized TPX2 is required for nearly all Aurora A activation and for full p53 synthesis and phosphorylation in vivo during oocyte maturation. In vitro, phosphorylation mediated by Aurora A targets serines 129 and 190 within the DNA binding domain of p53. Glutathione S-transferase pull-down studies indicate that the interaction occurs via the p53 transactivation domain and the Aurora A catalytic domain around the T-loop. Our studies suggest that targeting of TPX2 might be an effective strategy for specifically inhibiting the phosphorylation of Aurora A substrates, including p53.Aurora A is an oncogenic protein kinase that is active in mitosis and plays important roles in spindle assembly and centrosome function (1). Overexpression of either human or Xenopus Aurora A transforms mammalian cells, but only when the p53 pathway is altered (24). Aurora A is localized on centrosomes during mitosis, and overexpression of the protein leads to centrosome amplification and aneuploidy (2, 3, 5, 6), two likely contributors to genomic instability (7, 8). Because of its oncogenic potential and amplification in human tumors, considerable attention has been focused on the mechanism of Aurora A activation in mitosis. Evidence from several laboratories indicates that activation occurs as a result of phosphorylation of a threonine residue in the T-loop of the kinase (4, 9, 10). Purification of Aurora A-activating activity from M phase Xenopus egg extracts led to an apparent activation mechanism in which autophosphorylation at the T-loop is stimulated by binding of the targeting protein for Xklp2 (TPX2) (1114). On the other hand, it has been shown that Aurora A activity can be inhibited by interaction with several proteins, including PP1 (protein phosphatase 1), AIP (Aurora A kinase-interacting protein), and, more recently, p53 (9, 1517).p53 is a well known tumor suppressor able to drive cell cycle arrest, apoptosis, or senescence when DNA is damaged or cell integrity is threatened (18, 19). In human cancers, the p53 gene is frequently deleted or mutated, leading to inactivation of p53 functions (20). p53 protein is almost undetectable in “normal cells,” mainly due to its instability. Indeed, during a normal cell cycle, p53 associates with Mdm2 in the nucleus and thereafter undergoes nuclear exclusion, allowing its ubiquitination and subsequent degradation (21). In cells under stress, p53 is stabilized through the disruption of its interaction with Mdm2 (21), leading to p53 accumulation in the nucleus and triggering different responses, as described above.Although p53 has mostly been characterized as a nuclear protein, it has also been shown to localize on centrosomes (2224) and regulate centrosome duplication (23, 24). Centrosomes are believed to act as scaffolds that concentrate many regulatory molecules involved in signal transduction, including multiple protein kinases (25). Thus, centrosomal localization of p53 might be important for its own regulation by phosphorylation/dephosphorylation, and one of its regulators could be the mitotic kinase Aurora A. Indeed, phenotypes associated with the misexpression of these two proteins are very similar. For example, overexpression of Aurora A kinase leads to centrosome amplification, aneuploidy, and tumorigenesis, and the same effects are often observed after down-regulation of p53 transactivation activity or deletion/mutation of its gene (26, 27).Several recent studies performed in mammalian models show interplay between p53 and Aurora A, with each protein having the ability to inhibit the other, depending on the stage of the cell cycle and the stress level of the cell (17, 28, 29). These studies reported that p53 is a substrate of Aurora A, and serines 215 and 315 were demonstrated to be the two major Aurora A phosphorylation sites in human p53 in vitro and in vivo. Phosphorylation of Ser-215 within the DNA binding domain of human p53 inhibited both p53 DNA binding and transactivation activities (29). Recently, our group showed that Xenopus p53 is able to inhibit Aurora A kinase activity in vitro, but this inhibitory effect can be suppressed by prior binding of Aurora A to TPX2 (9). Contrary to somatic cells, where p53 is nuclear, unstable, and expressed at a very low level, p53 is highly expressed in the cytoplasm of Xenopus oocytes and stable until later stages of development (30, 31). The high concentration of both p53 and Aurora A in the oocyte provided a suitable basis for investigating p53-Aurora A interaction and also evaluating Xenopus p53 as a substrate of Aurora A.  相似文献   

19.
The Notch receptor is critical for proper development where it orchestrates numerous cell fate decisions. The Fringe family of β1,3-N-acetylglucosaminyltransferases are regulators of this pathway. Fringe enzymes add N-acetylglucosamine to O-linked fucose on the epidermal growth factor repeats of Notch. Here we have analyzed the reaction catalyzed by Lunatic Fringe (Lfng) in detail. A mutagenesis strategy for Lfng was guided by a multiple sequence alignment of Fringe proteins and solutions from docking an epidermal growth factor-like O-fucose acceptor substrate onto a homology model of Lfng. We targeted three main areas as follows: residues that could help resolve where the fucose binds, residues in two conserved loops not observed in the published structure of Manic Fringe, and residues predicted to be involved in UDP-N-acetylglucosamine (UDP-GlcNAc) donor specificity. We utilized a kinetic analysis of mutant enzyme activity toward the small molecule acceptor substrate 4-nitrophenyl-α-l-fucopyranoside to judge their effect on Lfng activity. Our results support the positioning of O-fucose in a specific orientation to the catalytic residue. We also found evidence that one loop closes off the active site coincident with, or subsequent to, substrate binding. We propose a mechanism whereby the ordering of this short loop may alter the conformation of the catalytic aspartate. Finally, we identify several residues near the UDP-GlcNAc-binding site, which are specifically permissive toward UDP-GlcNAc utilization.Defects in Notch signaling have been implicated in numerous human diseases, including multiple sclerosis (1), several forms of cancer (2-4), cerebral autosomal dominant arteriopathy with sub-cortical infarcts and leukoencephalopathy (5), and spondylocostal dysostosis (SCD)3 (6-8). The transmembrane Notch signaling receptor is activated by members of the DSL (Delta, Serrate, Lag2) family of ligands (9, 10). In the endoplasmic reticulum, O-linked fucose glycans are added to the epidermal growth factor-like (EGF) repeats of the Notch extracellular domain by protein O-fucosyltransferase 1 (11-13). These O-fucose monosaccharides can be elongated in the Golgi apparatus by three highly conserved β1,3-N-acetylglucosaminyltransferases of the Fringe family (Lunatic (Lfng), Manic (Mfng), and Radical Fringe (Rfng) in mammals) (14-16). The formation of this GlcNAc-β1,3-Fuc-α1, O-serine/threonine disaccharide is necessary and sufficient for subsequent elongation to a tetrasaccharide (15, 19), although elongation past the disaccharide in Drosophila is not yet clear (20, 21). Elongation of O-fucose by Fringe is known to potentiate Notch signaling from Delta ligands and inhibit signaling from Serrate ligands (22). Delta ligands are termed Delta-like (Delta-like1, -2, and -4) in mammals, and the homologs of Serrate are known as Jagged (Jagged1 and -2) in mammals. The effects of Fringe on Drosophila Notch can be recapitulated in Notch ligand in vitro binding assays using purified components, suggesting that the elongation of O-fucose by Fringe alters the binding of Notch to its ligands (21). Although Fringe also appears to alter Notch-ligand interactions in mammals, the effects of elongation of the glycan past the O-fucose monosaccharide is more complicated and appears to be cell type-, receptor-, and ligand-dependent (for a recent review see Ref. 23).The Fringe enzymes catalyze the transfer of GlcNAc from the donor substrate UDP-α-GlcNAc to the acceptor fucose, forming the GlcNAc-β1,3-Fuc disaccharide (14-16). They belong to the GT-A-fold of inverting glycosyltransferases, which includes N-acetylglucosaminyltransferase I and β1,4-galactosyltransferase I (17, 18). The mechanism is presumed to proceed through the abstraction of a proton from the acceptor substrate by a catalytic base (Asp or Glu) in the active site. This creates a nucleophile that attacks the anomeric carbon of the nucleotide-sugar donor, inverting its configuration from α (on the nucleotide sugar) to β (in the product) (24, 25). The enzyme then releases the acceptor substrate modified with a disaccharide and UDP. The Mfng structure (26) leaves little doubt as to the identity of the catalytic residue, which in all likelihood is aspartate 289 in mouse Lfng (we will use numbering for mouse Lunatic Fringe throughout, unless otherwise stated). The structure of Mfng with UDP-GlcNAc soaked into the crystals (26) showed density only for the UDP portion of the nucleotide-sugar donor and no density for two loops flanking either side of the active site. The presence of flexible loops that become ordered upon substrate binding is a common observation with glycosyltransferases in the GT-A fold family (18, 25). Density for the entire donor was observed in the structure of rabbit N-acetylglucosaminyltransferase I (27). In this case, ordering of a previously disordered loop upon UDP-GlcNAc binding may have contributed to increased stability of the donor. In the case of bovine β1,4-galactosyltransferase I, a section of flexible random coil from the apo-structure was observed to change its conformation to α-helical upon donor substrate binding (28). Both loops in Lfng are highly conserved, and we have mutated a number of residues in each to test the hypothesis that they interact with the substrates. The mutagenesis strategy was also guided by docking of an EGF-O-fucose acceptor substrate into the active site of the Lfng model as well as comparison of the Lfng model with a homology model of the β1,3-glucosyltransferase (β3GlcT) that modifies O-fucose on thrombospondin type 1 repeats (29, 30). The β3GlcT is predicted to be a GT-A fold enzyme related to the Fringe family (17, 18, 29).  相似文献   

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