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.
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   

2.
3.
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   

4.
Macroautophagy is a vacuolar lysosomal catabolic pathway that is stimulated during periods of nutrient starvation to preserve cell integrity. Ceramide is a bioactive sphingolipid associated with a large range of cell processes. Here we show that short-chain ceramides (C2-ceramide and C6-ceramide) and stimulation of the de novo ceramide synthesis by tamoxifen induce the dissociation of the complex formed between the autophagy protein Beclin 1 and the anti-apoptotic protein Bcl-2. This dissociation is required for macroautophagy to be induced either in response to ceramide or to starvation. Three potential phosphorylation sites, Thr69, Ser70, and Ser87, located in the non-structural N-terminal loop of Bcl-2, play major roles in the dissociation of Bcl-2 from Beclin 1. We further show that activation of c-Jun N-terminal protein kinase 1 by ceramide is required both to phosphorylate Bcl-2 and to stimulate macroautophagy. These findings reveal a new aspect of sphingolipid signaling in up-regulating a major cell process involved in cell adaptation to stress.Macroautophagy (referred to below as “autophagy”) is a vacuolar, lysosomal degradation pathway for cytoplasmic constituents that is conserved in eukaryotic cells (13). Autophagy is initiated by the formation of a multimembrane-bound autophagosome that engulfs cytoplasmic proteins and organelles. The last stage in the process results in fusion with the lysosomal compartments, where the autophagic cargo undergoes degradation. Basal autophagy is important in controlling the quality of the cytoplasm by removing damaged organelles and protein aggregates. Inhibition of basal autophagy in the brain is deleterious, and leads to neurodegeneration in mouse models (4, 5). Stimulation of autophagy during periods of nutrient starvation is a physiological response present at birth and has been shown to provide energy in various tissues of newborn pups (6). In cultured cells, starvation-induced autophagy is an autonomous cell survival mechanism, which provides nutrients to maintain a metabolic rate and level of ATP compatible with cell survival (7). In addition, starvation-induced autophagy blocks the induction of apoptosis (8). In other contexts, such as drug treatment and a hypoxic environment, autophagy has also been shown to be cytoprotective in cancer cells (9, 10). However, autophagy is also part of cell death pathways in certain situations (11). Autophagy can be a player in apoptosis-independent type-2 cell death (type-1 cell death is apoptosis), also known as autophagic cell death. This situation has been shown to occur when the apoptotic machinery is crippled in mammalian cells (12, 13). Autophagy can also be part of the apoptotic program, for instance in tumor necrosis factor-α-induced cell death when NF-κB is inhibited (14), or in human immunodeficiency virus envelope-mediated cell death in bystander naive CD4 T cells (15). Moreover autophagy has recently been shown to be required for the externalization of phosphatidylserine, the eat-me signal for phagocytic cells, at the surface of apoptotic cells (16).The complex relationship between autophagy and apoptosis reflects the intertwined regulation of these processes (17, 18). Many signaling pathways involved in the regulation of autophagy also regulate apoptosis. This intertwining has recently been shown to occur at the level of the molecular machinery of autophagy. In fact the anti-apoptotic protein Bcl-2 has been shown to inhibit starvation-induced autophagy by interacting with the autophagy protein Beclin 1 (19). Beclin 1 is one of the Atg proteins conserved from yeast to humans (it is the mammalian orthologue of yeast Atg6) and is involved in autophagosome formation (20). Beclin 1 is a platform protein that interacts with several different partners, including hVps34 (class III phosphatidylinositol 3-kinase), which is responsible for the synthesis of phosphatidylinositol 3-phosphate. The production of this lipid is important for events associated with the nucleation of the isolation membrane before it elongates and closes to form autophagosomes in response to other Atg proteins, including the Atg12 and LC32 (microtubule-associated protein light chain 3 is the mammalian orthologue of the yeast Atg8) ubiquitin-like conjugation systems (3, 21). Various partners associated with the Beclin 1 complex modulate the activity of hVps34. For instance, Bcl-2 inhibits the activity of this enzyme, whereas UVRAG, Ambra-1, and Bif-1 all up-regulate it (22, 23).In view of the intertwining between autophagy and apoptosis, it is noteworthy that Beclin 1 belongs to the BH3-only family of proteins (2426). However, and unlike most of the proteins in this family, Beclin 1 is not able to trigger apoptosis when its expression is forced in cells (27). A BH3-mimetic drug, ABT-737, is able to dissociate the Beclin 1-Bcl-2 complex, and to trigger autophagy by mirroring the effect of starvation (25).The sphingolipids constitute a family of bioactive lipids (2832) of which several members, such as ceramide and sphingosine 1-phosphate, are signaling molecules. These molecules constitute a “sphingolipid rheostat” that determines the fate of the cell, because in many settings ceramide is pro-apoptotic and sphingosine 1-phosphate mitigates this apoptotic effect (31, 32). However, ceramide is also engaged in a wide variety of other cell processes, such as the formation of exosomes (33), differentiation, cell proliferation, and senescence (34). Recently we showed that both ceramide and sphingosine 1-phosphate are able to stimulate autophagy (35, 36). It has also been shown that ceramide triggers autophagy in a large panel of mammalian cells (3739). However, elucidation of the mechanism by which ceramide stimulates autophagy is still in its infancy. We have previously demonstrated that ceramide induces autophagy in breast and colon cancer cells by inhibiting the Class I phosphatidylinositol 3-phosphate/mTOR signaling pathway, which plays a central role in inhibiting autophagy (36). Inhibition of mTOR is another hallmark of starvation-induced autophagy (17). This finding led us to investigate the effect of ceramide on the Beclin 1-Bcl-2 complex. The results presented here show that ceramide is more potent than starvation in dissociating the Beclin 1-Bcl-2 complex (see Ref. 40). This dissociation is dependent on three phosphorylation sites (Thr69, Ser70, and Ser87) located in a non-structural loop of Bcl-2. Ceramide induces the c-Jun N-terminal kinase 1-dependent phosphorylation of Bcl-2. Expression of a dominant negative form of JNK1 blocks Bcl-2 phosphorylation, and thus the induction of autophagy by ceramide. These findings help to explain how autophagy is regulated by a major lipid second messenger.  相似文献   

5.
The causative agent of Legionnaires disease, Legionella pneumophila, forms a replicative vacuole in phagocytes by means of the intracellular multiplication/defective organelle trafficking (Icm/Dot) type IV secretion system and translocated effector proteins, some of which subvert host GTP and phosphoinositide (PI) metabolism. The Icm/Dot substrate SidC anchors to the membrane of Legionella-containing vacuoles (LCVs) by specifically binding to phosphatidylinositol 4-phosphate (PtdIns(4)P). Using a nonbiased screen for novel L. pneumophila PI-binding proteins, we identified the Rab1 guanine nucleotide exchange factor (GEF) SidM/DrrA as the predominant PtdIns(4)P-binding protein. Purified SidM specifically and directly bound to PtdIns(4)P, whereas the SidM-interacting Icm/Dot substrate LidA preferentially bound PtdIns(3)P but also PtdIns(4)P, and the L. pneumophila Arf1 GEF RalF did not bind to any PIs. The PtdIns(4)P-binding domain of SidM was mapped to the 12-kDa C-terminal sequence, termed “P4M” (PtdIns4P binding of SidM/DrrA). The isolated P4M domain is largely helical and displayed higher PtdIns(4)P binding activity in the context of the α-helical, monomeric full-length protein. SidM constructs containing P4M were translocated by Icm/Dot-proficient L. pneumophila and localized to the LCV membrane, indicating that SidM anchors to PtdIns(4)P on LCVs via its P4M domain. An L. pneumophila ΔsidM mutant strain displayed significantly higher amounts of SidC on LCVs, suggesting that SidM and SidC compete for limiting amounts of PtdIns(4)P on the vacuole. Finally, RNA interference revealed that PtdIns(4)P on LCVs is specifically formed by host PtdIns 4-kinase IIIβ. Thus, L. pneumophila exploits PtdIns(4)P produced by PtdIns 4-kinase IIIβ to anchor the effectors SidC and SidM to LCVs.The Gram-negative pathogen Legionella pneumophila is the causative agent of Legionnaires disease, but it evolved as a parasite of various species of environmental predatory protozoa, including the social amoeba Dictyostelium discoideum (1, 2). The human disease is linked to the inhalation of contaminated aerosols, followed by replication in alveolar macrophages. To accommodate the transfer between host cells, L. pneumophila alternates between replicative and transmissive phases, the regulation of which includes an apparent quorum-sensing system (35).In macrophages and amoebae, L. pneumophila forms a replicative compartment, the Legionella-containing vacuole (LCV).3 LCVs avoid fusion with lysosomes (6), intercept vesicular traffic at endoplasmic reticulum (ER) exit sites (7), and fuse with the ER (810). The uptake of L. pneumophila and formation of LCVs in macrophages and amoebae depends on the Icm/Dot type IV secretion system (T4SS) (1114). Although more than 100 Icm/Dot substrates (“effector” proteins) have been identified to date, only few are functionally characterized, including effectors that interfere with host cell signal transduction, vesicle trafficking, or apoptotic pathways (1518).Two Icm/Dot-translocated substrates, SidM/DrrA (19, 20) and RalF (21), have been characterized as guanine nucleotide exchange factors (GEFs) for the Rho subfamily of small GTPases. These bacterial GEFs are recruited to and activate their targets on LCVs. Small GTPases of the Rho subfamily are involved in many eukaryotic signal transduction pathways and in actin cytoskeleton regulation (22). Inactive Rho GTPases bind GDP and a guanine nucleotide dissociation inhibitor (GDI). The GTPases are activated by removal of the GDI and the exchange of GDP with GTP by GEFs, which promotes the interaction with downstream effector proteins, such as protein or lipid kinases and various adaptor proteins. The cycle is closed by hydrolysis of the bound GTP, which is mediated by GTPase-activating proteins.SidM is a GEF for Rab1, which is essential for ER to Golgi vesicle transport, and additionally, SidM acts as a GDI displacement factor (GDF) to activate Rab1 (23, 24). The function of SidM is assisted by the Icm/Dot substrate LidA, which also localizes to LCVs. LidA preferentially binds to activated Rab1, thus supporting the recruitment of early secretory vesicles by SidM (19, 20, 23, 25, 26). Another Icm/Dot substrate, LepB (27), contributes to Rab1-mediated membrane cycling by inactivating Rab1 through its GTPase-activating protein function, thus acting as an antagonist of SidM (24).The Icm/Dot substrate RalF recruits and activates the small GTPase ADP-ribosylation factor 1 (Arf1), which is involved in retrograde vesicle transport from Golgi to ER (21). Dominant negative Arf1 (7, 28) or knockdown of Arf1 by RNA interference (29) impairs the formation of LCVs, as well as the recruitment of the Icm/Dot substrate SidC to the LCV (30).SidC and its paralogue SdcA localize to the LCV membrane (31), where the proteins specifically bind to the host cell lipid phosphatidylinositol 4-phosphate (PtdIns(4)P) (32, 33). Phosphoinositides (PIs) regulate eukaryotic receptor-mediated signal transduction, actin remodeling, and membrane dynamics (34, 35). PtdIns(4)P is present on the cytoplasmic membrane, but localizes preferentially to the trans-Golgi network (TGN), where this PI is produced by an Arf-dependent recruitment of PtdIns(4)P kinase IIIβ (PI4K IIIβ) (36) to promote trafficking along the secretory pathway. Recently, PtdIns(4)P was found to also mediate the export of early secretory vesicles from ER exit sites (37). At present, the L. pneumophila effector proteins that mediate exploitation of host PI signaling remain ill defined.In a nonbiased screen for L. pneumophila PI-binding proteins using different PIs coupled to agarose beads, we identified SidM as a major PtdIns(4)P-binding effector. We mapped its PtdIns(4)P binding activity to a novel P4M domain within a 12-kDa C-terminal sequence. SidM constructs, including the P4M domain, were found to be translocated and bind the LCV membrane, where the levels of PtdIns(4)P are controlled by PI4K IIIβ.  相似文献   

6.
7.
Cell-cell adhesion is a dynamic process that can activate multiple signaling pathways. These signaling pathways can be regulated through reversible tyrosine phosphorylation events. The level of tyrosine phosphorylation of junctional proteins reflects the balance between protein-tyrosine kinase and protein-tyrosine phosphatase activity. The receptor-tyrosine phosphatase DEP-1 (CD148/PTP-η) has been implicated in cell growth and differentiation as well as in regulating phosphorylation of junctional proteins. However, the role of DEP-1 in regulating tight junction phosphorylation and the integrity of cell-cell junctions is still under investigation. In this study, we used a catalytically dead substrate-trapping mutant of DEP-1 to identify potential substrates at cell-cell junctions. We have shown that in epithelial cells the trapping mutant of DEP-1 interacts with the tight junction proteins occludin and ZO-1 in a tyrosine phosphorylation-dependent manner. In contrast, PTP-PEST, Shp2, and PTPμ did not interact with these proteins, suggesting that the interaction of DEP-1 with occludin and ZO-1 is specific. In addition, occludin and ZO-1 were dephosphorylated by DEP-1 but not these other phosphatases in vitro. Overexpression of DEP-1 increased barrier function as measured by transepithelial electrical resistance and also reduced paracellular flux of fluorescein isothiocyanate-dextran following a calcium switch. Reduced DEP-1 expression by small interfering RNA had a small but significant increase in junction permeability. These data suggest that DEP-1 can modify the phosphorylation state of tight junction proteins and play a role in regulating permeability.Tight junctions are the most apical of junctions formed by epithelia and provide a regulated barrier to paracellular transport of ions, solutes, macromolecules, and even other cells. In addition, tight junctions act as a “fence” within the plane of the membrane, dividing the apical and basolateral domains of polarized epithelial cells. These junctions play an important role in the regulation of multiple cellular processes including cell differentiation, proliferation, and polarity (for reviews see Refs. 1 and 2). Functional tight junctions are characterized by the presence of membrane spanning proteins (claudins, occludin, and JAMs), which interact with cytoplasmic proteins (AF-6 and ZO-1, -2, -3), regulating assembly and maintenance of tight junctions. Occludin spans the membrane four times and was the first transmembrane component of the tight junction to be identified (3). It has two extracellular regions, an intracellular loop, as well as both an N- and C-terminal cytoplasmic tail (3). The C-terminal tail of occludin binds directly to the ZO family of proteins, which link the protein complex to the actin cytoskeleton (48). The long C-terminal domain is rich in serine, threonine, and tyrosine residues (9). In fact, several kinases and phosphatases interact with and modulate the phosphorylation state of tight junction proteins (1014). Serine and threonine phosphorylation of occludin is abundant in epithelia with intact junctions, whereas tyrosine phosphorylation is undetectable (15). However, tyrosine phosphorylation of occludin is associated with a decrease in transepithelial electrical resistance (TER)2 (16, 17) and loss of protein localization at the tight junction (18). Increases in tyrosine phosphorylation of occludin and ZO-1 result in the dissociation of the occludin-ZO-1 complex and reduces the localization at the tight junction of these proteins (12, 19). These data suggest that the phosphorylation state of tight junction proteins can affect the integrity of the tight junction complex and therefore the integrity of the tight junction itself. Both serine and threonine kinases and phosphatases bind to and act on TJ proteins (reviewed in Ref. 20). Previous studies have identified that c-Src and c-Yes are protein-tyrosine kinases, which act on the TJ, however, to date no protein-tyrosine phosphatases have been specifically characterized as acting on TJ proteins (1113, 21).Similarly, the other major junction of epithelia, the adherens junction (AJ), is also regulated by tyrosine phosphorylation. Increased tyrosine phosphorylation of the AJ decreases the stability of the cadherin-catenin complex, disrupting the association with the cytoskeleton and reducing junctional integrity (2224). Therefore, these studies suggest that maintenance of junctional integrity for both the TJ and AJ is regulated in part by reversible tyrosine phosphorylation that results from a competing balance of protein-tyrosine kinase and protein-tyrosine phosphatase (PTP) activity. Several PTPs have been localized to AJs and shown to bind components of the cadherin-catenin complex. The PTPs in AJs include receptor-PTPs (PTPμ, DEP-1, and vascular endothelial-PTP), as well as cytosolic PTPs (PTP1B and Shp-2) (2530). The high concentration of PTPs at cell-cell junctions indicates the importance of maintaining low levels of tyrosine phosphorylation except when the junctions need to be remodeled or disassembled.DEP-1 (density-enhanced phosphatase-1) is a receptor PTP that was first cloned from a human cDNA library and named based on the observation that its expression was elevated with increasing cell density (31). Also known as PTP-η, PTPRJ, and CD148, DEP-1 is comprised of an extracellular domain of eight fibronectin type III repeats, a transmembrane domain, and a single cytoplasmic catalytic domain. The protein is ubiquitously expressed (32), indicating its potential involvement in a large number of diverse signaling pathways. DEP-1 is involved in regulating the differentiation of epithelial cells (3336), as well as controlling cell growth and adhesion (33, 34). In addition, DEP-1 is able to attenuate the cellular response to growth factors through the preferential dephosphorylation of several growth factor receptors, suggesting that DEP-1 can selectively dephosphorylate certain tyrosines to more finely control signaling (3741).In addition to its role in proliferation and differentiation, DEP-1 localizes to areas of cell-cell adhesion in endothelial and epithelial cells, overlapping with the AJ marker vascular endothelial-cadherin in endothelia (42). Interaction with p120 catenin as well as other members of the catenin family also supports the hypothesis that DEP-1 plays a role in regulating AJ protein phosphorylation (27, 40). In the current study, we investigated whether adjacent tight junction proteins are also substrates of DEP-1. We now demonstrate that the substrate-trapping mutant of DEP-1 interacts with the tight junction proteins occludin and ZO-1. The association of DEP-1 with occludin and ZO-1 is specific to DEP-1 and not other phosphatases tested. In addition, DEP-1 is able to dephosphorylate occludin and ZO-1 indicating that these tight junction proteins are substrates of DEP-1. Furthermore, increased expression of DEP-1 enhances barrier function as junctions reform following a calcium switch and loss of DEP-1 levels increased the permeability of a stable epithelial monolayer. Together these results indicate that ZO-1 and occludin are substrates of DEP-1 and imply a role for DEP-1 in influencing the phosphorylation state of tight junction proteins and junction permeability.  相似文献   

8.
9.
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   

10.
The Tob/BTG family is a group of antiproliferative proteins containing two highly homologous regions, Box A and Box B. These proteins all associate with CCR4-associated factor 1 (Caf1), which belongs to the ribonuclease D (RNase D) family of deadenylases and is a component of the CCR4-Not deadenylase complex. Here we determined the crystal structure of the complex of the N-terminal region of Tob and human Caf1 (hCaf1). Tob exhibited a novel fold, whereas hCaf1 most closely resembled the catalytic domain of yeast Pop2 and human poly(A)-specific ribonuclease. Interestingly, the association of hCaf1 was mediated by both Box A and Box B of Tob. Cell growth assays using both wild-type and mutant proteins revealed that deadenylase activity of Caf1 is not critical but complex formation is crucial to cell growth inhibition. Caf1 tethers Tob to the CCR4-Not deadenylase complex, and thereby Tob gathers several factors at its C-terminal region, such as poly(A)-binding proteins, to exert antiproliferative activity.The Tob/BTG family (also called the APRO family) is a group of antiproliferative proteins (1, 2) consisting of Tob (3), Tob2 (4), BTG1 (5), BTG2/Tis21/PC3 (6-8), PC3B (9), and ANA/BTG3 (10, 11) in mammalian cells, AF177464 in Drosophila, and FOG-3 in Caenorhabditis elegans (12). A recent genome project reported that the BTG/Tob family protein had already existed in Choanoflagellida Monosiga brevicollis MX1. The N-terminal region of the Tob/BTG family proteins is conserved and includes two highly homologous regions, Box A and Box B. The Tob/BTG family proteins are involved in cell cycle regulation in a variety of cells such as T lymphocytes, fibroblasts, epithelial cells, and germ cells. In Tob-deficient mice, the incidence of liver tumors is higher than in wild-type mice. Furthermore, because the levels of tob expression are often repressed in human lung cancers, suppression of its expression is thought to contribute to tumor progression (13).The antiproliferative activities of the Tob/BTG family proteins are due to their association with target proteins in cells. For example, Tob associates with SMAD family proteins and acts as a negative regulator of SMAD signaling. In osteoblasts, this negative regulation occurs via association with SMAD 1, 5, 6, and 8 (14, 15), and via association with SMAD 2 and 4 in anergic quiescent T cells (16). Tob/BTG family proteins also bind to protein arginine methyltransferase, which regulates chromatin assembly by histone methylation (17). Much evidence has been accumulated to suggest that CCR4-associated factor 1 (Caf1),2 also known as Cnot7 and involved in the CCR4-Not deadenylase complex, is a common binding partner of the Tob/BTG family proteins (4, 18-21). To reveal the functions of Caf1 in vivo, caf1-/- mice have been generated in two groups. Male caf1-deficient mice are infertile because of a malfunction of the testicular somatic cells that leads to a defect in spermatogenesis (22, 23). Genetic analysis of the nematode C. elegans also suggests that FOG3 (Tob orthologue) interacts with CCF1, the C. elegans homologue of Caf1, and that this interaction is essential for germ cells to initiate spermatogenesis (24).Mouse and human Caf1 (mCaf1 and hCaf1) were found as homologues of yeast Pop2, a component of the CCR4-Not complex (18, 25). Yeast Pop2 displays weak RNase activity but enhances the deadenylation of the poly(A) tail of mRNA by the CCR4-Not deadenylase complex (26-29). The primary structure of mammalian Caf1 is related to that of the ribonuclease D (RNase D) family, and all of the active site residues are well conserved (30). Indeed, both mCaf1 and hCaf1 have deadenylase activity (31-33).Although the relationship between cell cycle repression and poly(A) deadenylation is not well understood, mRNA degradation and synthesis are major events in maintaining the cell cycle (34). The mRNAs in a eukaryotic cell have a wide range of half-lives. Degradation of mRNA is initiated by shortening of the poly(A) tail. Thereafter, the 5′-cap structure is removed and the remaining portion of the mRNA is rapidly degraded. The degradation of eukaryotic mRNAs is regulated precisely at each stage of the cell cycle. Tob was reported to associate with inducible poly(A)-binding protein (iPABP) and to abrogate the translation of interleukin-2 mRNA in vitro (35). Recent reports also showed that Tob and BTG2 interact with the CCR4-Not deadenylase complex using the Tob/BTG2 domain and the cytoplasmic poly(A)-binding protein (PABPC1) using the C-terminal tail and enhanced mRNA degradation (36-38).To help elucidate the relationship between the antiproliferative activity of Tob and the degradation of the poly(A) tail, we determined the crystal structure of the Tob-hCaf1 complex. We found that hCaf1 has a structure similar to yeast Pop2 and human PARN of deadenylases, exonuclease I, and the Klenow fragment of DNA polymerase I from Escherichia coli. In contrast, Tob has a novel structure. Specifically, Box A and Box B mediate the interaction between Tob and hCaf1. Cell growth assays using the wild and mutant proteins, together with the structural studies, revealed that the complex formation is crucial to cell growth inhibition.  相似文献   

11.
12.
Shugoshin 1 (Sgo1) protects centromeric sister-chromatid cohesion in early mitosis and, thus, prevents premature sister-chromatid separation. The protein level of Sgo1 is regulated during the cell cycle; it peaks in mitosis and is down-regulated in G1/S. Here we show that Sgo1 is degraded during the exit from mitosis, and its degradation depends on the anaphase-promoting complex/cyclosome (APC/C). Overexpression of Cdh1 reduces the protein levels of ectopically expressed Sgo1 in human cells. Sgo1 is ubiquitinated by APC/C bound to Cdh1 (APC/CCdh1) in vitro. We have further identified two functional degradation motifs in Sgo1; that is, a KEN (Lys-Glu-Asn) box and a destruction box (D box). Although removal of either motif is not sufficient to stabilize Sgo1, Sgo1 with both KEN box and D box deleted is stable in cells. Surprisingly, mitosis progresses normally in the presence of non-degradable Sgo1, indicating that degradation of Sgo1 is not required for sister-chromatid separation or mitotic exit. Finally, we show that the spindle checkpoint kinase Bub1 contributes to the maintenance of Sgo1 steady-state protein levels in an APC/C-independent mechanism.Loss of sister-chromatid cohesion triggers chromosome segregation in mitosis and occurs in two steps in vertebrate cells (1-3). In prophase, cohesin is phosphorylated by mitotic kinases including Plk1 and removed from chromosome arms (1, 4). Then, cleavage of centromeric cohesin by separase takes place at the metaphase-to-anaphase transition to allow sister-chromatid separation (5). The shugoshin (Sgo) family of proteins plays an important role in the protection of centromeric cohesion (6, 7). Human cells depleted of Sgo1 by RNAi undergo massive chromosome missegregation (8-11). In cells with compromised Sgo1 function, centromeric cohesin is improperly phosphorylated and removed (4, 11), resulting in premature sister-chromatid separation. It has been shown recently that Sgo1 collaborates with PP2A to counteract the action of Plk1 and other mitotic kinases and to protect centromeric cohesin from premature removal (12-14). In addition, Sgo1 has also been shown to promote stable kinetochore-microtubule attachment and sense tension across sister kinetochores (8, 15). Thus, Sgo1 is crucial for mitotic progression and chromosome segregation.Orderly progression through mitosis is regulated by the anaphase-promoting complex/cyclosome (APC/C),2 a large multiprotein ubiquitin ligase that targets key mitotic regulators for destruction by the proteasome (16). APC/C selects substrates for ubiquitination by using the Cdc20 or Cdh1 activator proteins to recognize specific sequences called APC/C degrons within target proteins (17). Several APC/C degrons have been characterized, including the destruction box (D box) and the Lys-Glu-Asn box (KEN box) (18, 19). The D box, with the consensus amino acid sequence of RXXLXXXN(X indicates any amino acid), are found in many APC/C substrates, including mitotic cyclins and are essential for their ubiquitin-mediated destruction. The KEN box, which contains a consensus KEN motif, is also found in several APC/C substrates and is preferentially but not exclusively recognized by APC/CCdh1. When APC/C is active, it directs progression through and exit from mitosis by catalyzing the ubiquitination and timely destruction of mitotic regulators, including cyclin A, cyclin B, and the separase inhibitor securin (16). The APC/C activity needs to be tightly controlled to prevent unscheduled substrate degradation. An important mechanism for APC/C regulation is the spindle checkpoint, which prevents the activation of APC/C and destruction of its substrates in response to kinetochores that have not properly attached to the mitotic spindle (20).Recent evidence shows that Sgo1 is a substrate of APC/C, and its protein levels oscillate during the cell cycle (8, 9). In this article we study the degradation of Sgo1 in human cells. We show that Sgo1 is degraded during mitotic exit, and this degradation depends on APC/CCdh1. We further show that both KEN and D boxes are required for Sgo1 degradation in vivo and ubiquitination in vitro. Removal of these motifs stabilizes Sgo1 in vivo. The prolonged presence of stable Sgo1 protein in human cells does not change the kinetics of chromosome segregation and mitotic exit. Therefore, a timely scheduled degradation of Sgo1 takes place but is not required for mitotic exit. Finally, we show that Bub1 regulates Sgo1 protein levels through a mechanism that does not involve APC/C-mediated degradation.  相似文献   

13.
It is known that platelet-activating factor (PAF) induces severe endothelial barrier leakiness, but the signaling mechanisms remain unclear. Here, using a wide range of biochemical and morphological approaches applied in both mouse models and cultured endothelial cells, we addressed the mechanisms of PAF-induced disruption of interendothelial junctions (IEJs) and of increased endothelial permeability. The formation of interendothelial gaps filled with filopodia and lamellipodia is the cellular event responsible for the disruption of endothelial barrier. We observed that PAF ligation of its receptor induced the activation of the Rho GTPase Rac1. Following PAF exposure, both Rac1 and its guanine nucleotide exchange factor Tiam1 were found associated with a membrane fraction from which they co-immunoprecipitated with PAF receptor. In the same time frame with Tiam1-Rac1 translocation, the junctional proteins ZO-1 and VE-cadherin were relocated from the IEJs, and formation of numerous interendothelial gaps was recorded. Notably, the response was independent of myosin light chain phosphorylation and thus distinct from other mediators, such as histamine and thrombin. The changes in actin status are driven by the PAF-induced localized actin polymerization as a consequence of Rac1 translocation and activation. Tiam1 was required for the activation of Rac1, actin polymerization, relocation of junctional associated proteins, and disruption of IEJs. Thus, PAF-induced IEJ disruption and increased endothelial permeability requires the activation of a Tiam1-Rac1 signaling module, suggesting a novel therapeutic target against increased vascular permeability associated with inflammatory diseases.The endothelial barrier is made up of endothelial cells (ECs)4 connected to each other by interendothelial junctions (IEJs) consisting of protein complexes organized as tight junctions (TJs) and adherens junctions (AJs). In addition, the focal adhesion complex located at the basal plasma membrane enables firm contact of ECs with the underlying basement membrane and also contributes to the barrier function (1-3). The glycocalyx, the endothelial monolayer, and the basement membrane all together constitute the vascular barrier.The structural integrity of the ECs along with their proper functionality are the two most important factors controlling the tightness of the endothelial barrier. Changes affecting these factors cause loss of barrier restrictiveness and leakiness. Therefore, defining and understanding the cellular and molecular mechanisms controlling these processes is of paramount importance. Increased width of IEJs in response to permeability-increasing mediators (4) regulates the magnitude of transendothelial exchange of fluid and solutes. Disruption of IEJs and the resultant barrier leakiness contribute to the genesis of diverse pathological conditions, such as inflammation (5), metastasis (6, 7), and uncontrolled angiogenesis (8, 9).Accumulated evidence demonstrated that IEJs changes are responsible for increased or decreased vascular permeability, and the generally accepted mechanism responsible for them was the myosin light chain (MLC)-mediated contraction of ECs (5, 10). However, published evidence showed that an increase in vascular permeability could be obtained without a direct involvement of any contractile mechanism (11-16).The main component of the vascular barrier, the ECs, has more than 10% of their total protein represented by actin (17), which under physiological salt concentrations subsists as monomers (G-actin) and assembled into filaments (F-actin). A large number of actin-interacting proteins may modulate the assembly, disassembly, and organization of G-actin and of actin filaments within a given cell type. Similar to the complexity of actin-interacting proteins found in other cell types, the ECs utilize their actin binding proteins to stabilize the endothelial monolayer in order to efficiently function as a selective barrier (11). In undisturbed ECs, the actin microfilaments are organized as different networks with distinctive functional and morphological characteristics: the peripheral filaments also known as peripheral dense band (PDB), the cytoplasmic fibers identified as stress fibers (SF), and the actin from the membrane cytoskeleton (18). The peripheral web, localized immediately under the membrane, is associated with (i) the luminal plasmalemma (on the apical side), (ii) the IEJ complexes on the lateral surfaces, and (iii) the focal adhesion complexes on the abluminal side (the basal part) of polarized ECs. The SF reside inside the endothelial cytoplasm and are believed to be directly connected with the plasmalemma proper on the luminal as well as on the abluminal side of the cell. As described, the endothelial actin cytoskeleton (specifically the SF) seems to be a stable structure helping the cells to remain flat under flow (19). It is also established that the actin fibers participate in correct localization of different junctional complexes while keeping them in place (20). However, it was suggested that the dynamic equilibrium between F- and G-actin might modulate the tightness of endothelial barrier in response to different challenges (13).Mediators effective at nanomolar concentrations or less that disrupt the endothelial barrier and increase vascular permeability include C2 toxin of Clostridium botulinum, vascular permeability factor, better known as vascular endothelial growth factor, and PAF (21). C2 toxin increases endothelial permeability by ribosylating monomeric G-actin at Arg-177 (22). This results in the impairment of actin polymerization (23), followed by rounding of ECs (16) and the disruption of junctional integrity. Vascular permeability factor was shown to open IEJs by redistribution of junctional proteins (24, 25) and by interfering with the equilibrium of actin pools (26). PAF (1-O-alkyl-2-acetyl-sn-glycero-3-phosphocoline), a naturally synthesized phospholipid is active at 10-10 m or less (27). PAF is synthesized by and acts on a variety of cell types, including platelets (28), neutrophils (29), monocytes (30), and ECs (31). PAF-mediated activation of ECs induced cell migration (32), angiogenesis (7), and vascular hyperpermeability (33) secondary to disassembly of IEJs (34). The effects of PAF on the endothelium are initiated through a G protein-coupled receptor (PAF-R) localized at the plasmalemma, in a large endosomal compartment inside the cell (34), and also in the nuclear membrane (35). In ECs, PAF-R was shown to signal through Gαq and downstream activation of phospholipase C isozymes (PLCβ3 and PLCγ1), and via cSrc (32, 36). Studies have shown that PAF challenge induced endothelial actin cytoskeletal rearrangement (37) and marked vascular leakiness (38); however, the signaling pathways have not been elucidated.Therefore, in the present study, we carried out a systematic analysis of PAF-induced morphological and biochemical changes of endothelial barrier in vivo and in cultured ECs. We found that the opening of endothelial barrier and the increased vascular leakiness induced by PAF are the result of a shift in actin pools without involvement of EC contraction, followed by a redistribution of tight junctional associated protein ZO-1 and adherens junctional protein VE-cadherin.  相似文献   

14.
15.
16.
17.
As obligate intracellular parasites, viruses exploit diverse cellular signaling machineries, including the mitogen-activated protein-kinase pathway, during their infections. We have demonstrated previously that the open reading frame 45 (ORF45) of Kaposi sarcoma-associated herpesvirus interacts with p90 ribosomal S6 kinases (RSKs) and strongly stimulates their kinase activities (Kuang, E., Tang, Q., Maul, G. G., and Zhu, F. (2008) J. Virol. 82 ,1838 -1850). Here, we define the mechanism by which ORF45 activates RSKs. We demonstrated that binding of ORF45 to RSK increases the association of extracellular signal-regulated kinase (ERK) with RSK, such that ORF45, RSK, and ERK formed high molecular mass protein complexes. We further demonstrated that the complexes shielded active pERK and pRSK from dephosphorylation. As a result, the complex-associated RSK and ERK were activated and sustained at high levels. Finally, we provide evidence that this mechanism contributes to the sustained activation of ERK and RSK in Kaposi sarcoma-associated herpesvirus lytic replication.The extracellular signal-regulated kinase (ERK)2 mitogen-activated protein kinase (MAPK) signaling pathway has been implicated in diverse cellular physiological processes including proliferation, survival, growth, differentiation, and motility (1-4) and is also exploited by a variety of viruses such as Kaposi sarcoma-associated herpesvirus (KSHV), human cytomegalovirus, human immunodeficiency virus, respiratory syncytial virus, hepatitis B virus, coxsackie, vaccinia, coronavirus, and influenza virus (5-17). The MAPK kinases relay the extracellular signaling through sequential phosphorylation to an array of cytoplasmic and nuclear substrates to elicit specific responses (1, 2, 18). Phosphorylation of MAPK is reversible. The kinetics of deactivation or duration of signaling dictates diverse biological outcomes (19, 20). For example, sustained but not transient activation of ERK signaling induces the differentiation of PC12 cells into sympathetic-like neurons and transformation of NIH3T3 cells (20-22). During viral infection, a unique biphasic ERK activation has been observed for some viruses (an early transient activation triggered by viral binding or entry and a late sustained activation correlated with viral gene expression), but the responsible viral factors and underlying mechanism for the sustained ERK activation remain largely unknown (5, 8, 13, 23).The p90 ribosomal S6 kinases (RSKs) are a family of serine/threonine kinases that lie at the terminus of the ERK pathway (1, 24-26). In mammals, four isoforms are known, RSK1 to RSK4. Each one has two catalytically functional kinase domains, the N-terminal kinase domain (NTKD) and C-terminal kinase domain (CTKD) as well as a linker region between the two. The NTKD is responsible for phosphorylation of exogenous substrates, and the CTKD and linker region regulate RSK activation (1, 24, 25). In quiescent cells ERK binds to the docking site in the C terminus of RSK (27-29). Upon mitogen stimulation, ERK is activated by its upstream MAPK/ERK kinase (MEK). The active ERK phosphorylates Thr-359/Ser-363 of RSK in the linker region (amino acid numbers refer to human RSK1) and Thr-573 in the CTKD activation loop. The activated CTKD then phosphorylates Ser-380 in the linker region, creating a docking site for 3-phosphoinositide-dependent protein kinase-1. The 3-phosphoinositide-dependent protein kinase-1 phosphorylates Ser-221 of RSK in the activation loop and activates the NTKD. The activated NTKD autophosphorylates the serine residue near the ERK docking site, causing a transient dissociation of active ERK from RSK (25, 26, 28). The stimulation of quiescent cells by a mitogen such as epidermal growth factor or a phorbol ester such as 12-O-tetradecanoylphorbol-13-acetate (TPA) usually results in a transient RSK activation that lasts less than 30 min. RSKs have been implicated in regulating cell survival, growth, and proliferation. Mutation or aberrant expression of RSK has been implicated in several human diseases including Coffin-Lowry syndrome and prostate and breast cancers (1, 24, 25, 30-32).KSHV is a human DNA tumor virus etiologically linked to Kaposi sarcoma, primary effusion lymphoma, and a subset of multicentric Castleman disease (33, 34). Infection and reactivation of KSHV activate multiple MAPK pathways (6, 12, 35). Noticeably, the ERK/RSK activation is sustained late during KSHV primary infection and reactivation from latency (5, 6, 12, 23), but the mechanism of the sustained ERK/RSK activation is unclear. Recently, we demonstrated that ORF45, an immediate early and also virion tegument protein of KSHV, interacts with RSK1 and RSK2 and strongly stimulates their kinase activities (23). We also demonstrated that the activation of RSK plays an essential role in KSHV lytic replication (23). In the present study we determined the mechanism of ORF45-induced sustained ERK/RSK activation. We found that ORF45 increases the association of RSK with ERK and protects them from dephosphorylation, causing sustained activation of both ERK and RSK.  相似文献   

18.
Occludin is phosphorylated on tyrosine residues during the oxidative stress-induced disruption of tight junction, and in vitro phosphorylation of occludin by c-Src attenuates its binding to ZO-1. In the present study mass spectrometric analyses of C-terminal domain of occludin identified Tyr-379 and Tyr-383 in chicken occludin as the phosphorylation sites, which are located in a highly conserved sequence of occludin, YETDYTT; Tyr-398 and Tyr-402 are the corresponding residues in human occludin. Deletion of YETDYTT motif abolished the c-Src-mediated phosphorylation of occludin and the regulation of ZO-1 binding. Y398A and Y402A mutations in human occludin also abolished the c-Src-mediated phosphorylation and regulation of ZO-1 binding. Y398D/Y402D mutation resulted in a dramatic reduction in ZO-1 binding even in the absence of c-Src. Similar to wild type occludin, its Y398A/Y402A mutant was localized at the plasma membrane and cell-cell contact sites in Rat-1 cells. However, Y398D/Y402D mutants of occludin failed to localize at the cell-cell contacts. Calcium-induced reassembly of Y398D/Y402D mutant occludin in Madin-Darby canine kidney cells was significantly delayed compared with that of wild type occludin or its T398A/T402A mutant. Furthermore, expression of Y398D/Y402D mutant of occludin sensitized MDCK cells for hydrogen peroxide-induced barrier disruption. This study reveals a unique motif in the occludin sequence that is involved in the regulation of ZO-1 binding by reversible phosphorylation of specific Tyr residues.Epithelial tight junctions (TJs)2 form a selective barrier to the diffusion of toxins, allergens, and pathogens from the external environment into the tissues in the gastrointestinal tract, lung, liver, and kidney (1). Disruption of TJs is associated with the gastrointestinal diseases such as inflammatory bowel disease, celiac disease, infectious enterocolitis, and colon cancer (24) as well as in diseases of lung and kidney (5, 6). Numerous inflammatory mediators such as tumor necrosis factor α, interferon γ, and oxidative stress (712) are known to disrupt the epithelial TJs and the barrier function. Several studies have indicated that hydrogen peroxide disrupts the TJs in intestinal epithelium by a tyrosine kinase-dependent mechanism (11, 12).Four types of integral proteins, occludin, claudins, junctional adhesion molecules, and tricellulin are associated with TJs. Occludin, claudins, and tricellulin are tetraspan proteins, and their extracellular domains interact with homotypic domains of the adjacent cells (1, 2, 13). The intracellular domains of these proteins interact with a variety of soluble proteins such as ZO-1, ZO-2, ZO-3, 7H6, cingulin, and symplekin (1423); this protein complex interacts with the perijunctional actomyosin ring. The interactions among TJ proteins are essential for the assembly and the maintenance of TJs. Therefore, regulation of the interactions among TJ proteins may regulate the TJ integrity. A significant body of evidence indicates that numerous signaling molecules are associated with the TJs. Protein kinases and protein phosphatases such as protein kinase Cζ (PKCζ), PKCι/λ (24), c-Src (25), c-Yes (26, 27), mitogen-activated protein kinase (28), PP2A, and PP1 (29) interact with TJs, indicating that TJs are dynamically regulated by intracellular signal transduction involving protein phosphorylation. Additionally, other signaling molecules such as calcium (30), phosphatidylinositol 3-kinase (31), Rho (32), and Rac (33) are involved in the regulation of TJs.Occludin, a ∼65-kDa protein, has been well characterized to be assembled into the TJs. Although occludin knock-out mice showed the formation of intact TJs in different epithelia (34), numerous studies have emphasized that it plays an important role in the regulation of TJ integrity. Occludin spans the membrane four times to form two extracellular loops and one intracellular loop, and the N-terminal and C-terminal domains hang into the intracellular compartment (3537). In epithelium with intact TJs, occludin is highly phosphorylated on Ser and Thr residues (38), whereas Tyr phosphorylation is undetectable. However, the disruption of TJs in Caco-2 cell monolayers by oxidative stress and acetaldehyde leads to Tyr phosphorylation of occludin; the tyrosine kinase inhibitors attenuate the disruption of TJs (39, 40). Furthermore, a previous in vitro study demonstrated that Tyr phosphorylation of the C-terminal domain of occludin leads to the loss of its interaction with ZO-1 and ZO-3 (25).In the present study we identified the Tyr residues in occludin that are phosphorylated by c-Src and determined their role in regulated interaction between occludin and ZO-1 and its assembly into the TJs. Results show that 1) Tyr-379 and Tyr-383 in chicken occludin and Tyr-398 and Tyr-402 in human occludin are the exclusive sites of phosphorylation by c-Src, and these Tyr residues are located in a highly conserved sequence of occludin, YET-DYTT, 2) deletion of YEDTYTT or point mutation of Tyr-398 and Tyr-402 in human occludin attenuates the phosphorylation-dependent regulation of ZO-1 binding, 3) Y398D/Y402D mutation of human occludin leads to loss of ZO-1 binding and prevents its translocation to the plasma membrane and cell-cell contact sites in Rat-1 cells, 4) Y398D/Y402D mutation of occludin delays its assembly into the intercellular junctions during the calcium-induced assembly of TJs, and 5) expression of Y398D/Y402D mutant occludin sensitizes cell monolayers for hydrogen peroxide-induced disruption of barrier function.  相似文献   

19.
The presence of extensive reciprocal conformational freedom between the catalytic and the hemopexin-like domains of full-length matrix metalloproteinase-1 (MMP-1) is demonstrated by NMR and small angle x-ray scattering experiments. This finding is discussed in relation to the essentiality of the hemopexin-like domain for the collagenolytic activity of MMP-1. The conformational freedom experienced by the present system, having the shortest linker between the two domains, when compared with similar findings on MMP-12 and MMP-9 having longer and the longest linker within the family, respectively, suggests this type of conformational freedom to be a general property of all MMPs.Matrix metalloproteinases (MMP)2 are extracellular hydrolytic enzymes involved in a variety of processes including connective tissue cleavage and remodeling (13). All 23 members of the family are able to cleave simple peptides derived from connective tissue components such as collagen, gelatin, elastin, etc. A subset of MMPs is able to hydrolyze more resistant polymeric substrates, such as cross-linked elastin, and partially degraded collagen forms, such as gelatin and type IV collagens (4). Intact triple helical type I–III collagen is only attacked by collagenases MMP-1, MMP-8, and MMP-13 and by MMP-2 and MMP-14 (512). Although the detailed mechanism of cleavage of single chain peptides by MMP has been largely elucidated (1319), little is known about the process of hydrolysis of triple helical collagen. In fact, triple helical collagen cannot be accommodated in the substrate-binding groove of the catalytic site of MMPs (9).All MMPs (but MMP-7) in their active form are constituted by a catalytic domain (CAT) and a hemopexin-like domain (HPX) (2022). The CAT domain contains two zinc ions and one to three calcium ions. One zinc ion is at the catalytic site and is responsible for the activity, whereas the other metal ions have structural roles. The isolated CAT domains retain full catalytic activity toward simple peptides and single chain polymeric substrates such as elastin, whereas hydrolysis of triple helical collagen also requires the presence of the HPX domain (9, 2325). It has been shown that the isolated CAT domain regains a small fraction of the activity of the full-length (FL) protein when high amounts of either inactivated full-length proteins or isolated HPX domains are added to the assay solution (9). Finally, it has been shown that the presence of the HPX domain alone alters the CD spectrum of triple helical collagen in a way that suggests its partial unwinding (26, 27). It is tempting to speculate that full-length collagenases attack collagen by first locally unwinding the triple helical structure with the help of the HPX domain and then cleaving the resulting, exposed, single filaments (9, 28).Until 2007, three-dimensional structures of full-length MMPs had been reported only for collagenase MMP-1 (2931) and gelatinase MMP-2 (32). The structures of the two proteins are very similar and show a compact arrangement of the two domains, which are connected by a short linker (14 and 20 amino acids, respectively). It is difficult to envisage that rigid and compact molecules of this type can interact with triple helical collagen in a way that can lead to first unwinding and then cleavage of individual filaments. It has been recently suggested that such concerted action could occur much more easily if the two domains could enjoy at least a partial conformational independence (9). Slight differences in the reciprocal orientation of the CAT and HPX domains of MMP-1 in the presence (29) and absence (30, 31) of the prodomain were indeed taken as a hint that the two domains could experience relative mobility (29).Two recent solution studies have shown that conformational independence is indeed occurring in gelatinase MMP-9 (33) and elastase MMP-12 (34), whereas the x-ray structure of the latter (34) is only slightly less compact than those of MMP-1 (2931) and MMP-2 (32). Among MMPs, MMP-9 features an exceptionally long linker (68 amino acid) (33, 35), which in fact constitutes a small domain by itself (the O-glycosylated domain) (33), and therefore, this inspiring observation can hardly be taken as evidence that conformational freedom is a general characteristic of the two-domain MMPs. MMP-12 features a much more normal 16-amino acid linker, thereby making more probable a general functional role for this conformational freedom (34). However, both MMP-9 and MMP-12 retain their full catalytic activity against their substrates even when deprived of the HPX domain (9). Therefore, the question remains of whether conformational freedom is also a required characteristic for those MMPs that are only active as full-length proteins, i.e. collagenases. Interestingly, the three collagenases (MMP-1, MMP-8, and MMP-13) have the shortest linker (14 amino acids) among all MMPs. Demonstrating or negating the presence of conformational freedom in one of these collagenases would therefore constitute a significant step forward to formulate mechanistic hypotheses on their collagenolytic activity.Our recent studies on MMP-12 in solution (34) have shown that a combination of NMR relaxation studies and small angle x-ray scattering (SAXS) is enough to show the presence and the extent of the relative conformational freedom of the two domains of MMPs. Here we apply the same strategy to full-length MMP-1 and show that sizable conformational freedom is indeed experienced even by this prototypical collagenase, although somewhat less pronounced than that observed for MMP-12.  相似文献   

20.
Hepatocellular carcinoma (HCC) is one of the most common and aggressive human malignancies. Recombinant tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) is a promising anti-tumor agent. However, many HCC cells show resistance to TRAIL-induced apoptosis. In this study, we showed that bortezomib, a proteasome inhibitor, overcame TRAIL resistance in HCC cells, including Huh-7, Hep3B, and Sk-Hep1. The combination of bortezomib and TRAIL restored the sensitivity of HCC cells to TRAIL-induced apoptosis. Comparing the molecular change in HCC cells treated with these agents, we found that down-regulation of phospho-Akt (P-Akt) played a key role in mediating TRAIL sensitization of bortezomib. The first evidence was that bortezomib down-regulated P-Akt in a dose- and time-dependent manner in TRAIL-treated HCC cells. Second, LY294002, a PI3K inhibitor, also sensitized resistant HCC cells to TRAIL-induced apoptosis. Third, knocking down Akt1 by small interference RNA also enhanced TRAIL-induced apoptosis in Huh-7 cells. Finally, ectopic expression of mutant Akt (constitutive active) in HCC cells abolished TRAIL sensitization effect of bortezomib. Moreover, okadaic acid, a protein phosphatase 2A (PP2A) inhibitor, reversed down-regulation of P-Akt in bortezomib-treated cells, and PP2A knockdown by small interference RNA also reduced apoptosis induced by the combination of TRAIL and bortezomib, indicating that PP2A may be important in mediating the effect of bortezomib on TRAIL sensitization. Together, bortezomib overcame TRAIL resistance at clinically achievable concentrations in hepatocellular carcinoma cells, and this effect is mediated at least partly via inhibition of the PI3K/Akt pathway.Hepatocellular carcinoma (HCC)2 is currently the fifth most common solid tumor worldwide and the fourth leading cause of cancer-related death. To date, surgery is still the only curative treatment but is only feasible in a small portion of patients (1). Drug treatment is the major therapy for patients with advanced stage disease. Unfortunately, the response rate to traditional chemotherapy for HCC patients is unsatisfactory (1). Novel pharmacological therapy is urgently needed for patients with advanced HCC. In this regard, the approval of sorafenib might open a new era of molecularly targeted therapy in the treatment of HCC patients.Tumor necrosis factor (TNF)-related apoptosis-inducing ligand (TRAIL), a type II transmembrane protein and a member of the TNF family, is a promising anti-tumor agent under clinical investigation (2). TRAIL functions by engaging its receptors expressed on the surface of target cells. Five receptors specific for TRAIL have been identified, including DR4/TRAIL-R1, DR5/TRAIL-R2, DcR1, DcR2, and osteoprotegerin. Among TRAIL receptors, only DR4 and DR5 contain an effective death domain that is essential to formation of death-inducing signaling complex (DISC), a critical step for TRAIL-induced apoptosis. Notably, the trimerization of the death domains recruits an adaptor molecule, Fas-associated protein with death domain (FADD), which subsequently recruits and activates caspase-8. In type I cells, activation of caspase-8 is sufficient to activate caspase-3 to induce apoptosis; however, in another type of cells (type II), the intrinsic mitochondrial pathway is essential for apoptosis characterized by cleavage of Bid and release of cytochrome c from mitochondria, which subsequently activates caspase-9 and caspase-3 (3).Although TRAIL induces apoptosis in malignant cells but sparing normal cells, some tumor cells are resistant to TRAIL-induced apoptosis. Mechanisms responsible for the resistance include receptors and intracellular resistance. Although the cell surface expression of DR4 or DR5 is absolutely required for TRAIL-induced apoptosis, tumor cells expressing these death receptors are not always sensitive to TRAIL due to intracellular mechanisms. For example, the cellular FLICE-inhibitory protein (c-FLIP), a homologue to caspase-8 but without protease activity, has been linked to TRAIL resistance in several studies (4, 5). In addition, inactivation of Bax, a proapoptotic Bcl-2 family protein, resulted in resistance to TRAIL in MMR-deficient tumors (6, 7), and reintroduction of Bax into Bax-deficient cells restored TRAIL sensitivity (8), indicating that the Bcl-2 family plays a critical role in intracellular mechanisms for resistance of TRAIL.Bortezomib, a proteasome inhibitor approved clinically for multiple myeloma and mantle cell lymphoma, has been investigated intensively for many types of cancer (9). Accumulating studies indicate that the combination of bortezomib and TRAIL overcomes the resistance to TRAIL in various types of cancer, including acute myeloid leukemia (4), lymphoma (1013), prostate (1417), colon (15, 18, 19), bladder (14, 16), renal cell carcinoma (20), thyroid (21), ovary (22), non-small cell lung (23, 24), sarcoma (25), and HCC (26, 27). Molecular targets responsible for the sensitizing effect of bortezomib on TRAIL-induced cell death include DR4 (14, 27), DR5 (14, 20, 2223, 28), c-FLIP (4, 11, 2123, 29), NF-κB (12, 24, 30), p21 (16, 21, 25), and p27 (25). In addition, Bcl-2 family also plays a role in the combinational effect of bortezomib and TRAIL, including Bcl-2 (10, 21), Bax (13, 22), Bak (27), Bcl-xL (21), Bik (18), and Bim (15).Recently, we have reported that Akt signaling is a major molecular determinant in bortezomib-induced apoptosis in HCC cells (31). In this study, we demonstrated that bortezomib overcame TRAIL resistance in HCC cells through inhibition of the PI3K/Akt pathway.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号