首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
3.
The lipid phosphatase PTEN functions as a tumor suppressor by dephosphorylating the D3 position of phosphoinositide-3,4,5-trisphosphate, thereby negatively regulating the phosphoinositide 3-kinase (PI3K)/AKT signaling pathway. In mammalian cells, PTEN exists either as a monomer or as a part of a >600-kDa complex (the PTEN-associated complex [PAC]). Previous studies suggest that the antagonism of PI3K/AKT signaling by PTEN may be mediated by a nonphosphorylated form of the protein resident within the multiprotein complex. Here we show that PTEN associates with p85, the regulatory subunit of PI3K. Using newly generated antibodies, we demonstrate that this PTEN-p85 association involves the unphosphorylated form of PTEN engaged within the PAC and also includes the p110β isoform of PI3K. The PTEN-p85 association is enhanced by trastuzumab treatment and linked to a decline in AKT phosphorylation in some ERBB2-amplified breast cancer cell lines. Together, these results suggest that integration of p85 into the PAC may provide a novel means of downregulating the PI3K/AKT pathway.The phosphoinositide 3-kinase (PI3K)/AKT signaling pathway regulates glucose/nutrient homeostasis and cell survival and plays a central role in both normal metabolism and cancer. The PTEN tumor suppressor gene (29, 30, 54) negatively regulates the PI3K/AKT pathway by dephosphorylating the D3 hydroxyl subunit of phosphoinositide-3,4,5-trisphosphate, a key membrane phosphatidylinositol generated by PI3K (34). PTEN undergoes genetic or epigenetic inactivation in many malignancies, including glioblastoma, melanoma, and endometrial, prostate, and breast cancers, among others (6, 13, 22, 23, 47, 49-51, 55, 68). Similarly, germ line mutations of PTEN are associated with the development of hamartomatous neoplasias such as Cowden disease and Bannayan-Zonana syndrome (17, 21, 41).The tumor suppressor function of PTEN undergoes dynamic regulation involving both C-terminal phosphorylation and protein-protein interactions. Phosphorylation of serine and threonine residues at the PTEN C-terminal tail, mediated by kinases such as CK2 and glycogen synthase kinase 3β, alters its conformational structure and association with PDZ domain-containing proteins and attenuates PTEN enzymatic activity (1, 11, 20, 32, 45, 61-63, 66, 67, 71). Conversely, PTEN function is promoted in large part through its stabilization in unphosphorylated form by incorporation into a high-molecular-weight protein complex (the PTEN-associated complex [PAC]) (66). We first demonstrated the existence of the PAC through gel filtration studies of rat liver extracts, which identified PTEN within a high-molecular-mass peak (>600 kDa), as well as a low-molecular-mass peak (40 to 100 kDa) in which PTEN is monomeric and phosphorylated (66). Subsequently, several PDZ domain-containing proteins were shown to interact with PTEN, including MAGI-1b, MAGI-2, MAGI-3, ghDLG, hMAST205, MSP58/MCRS1, NHERF1, and NHERF2, which mediate indirect binding with platelet-derived growth factor (PDGF) receptor β (25, 36, 42, 57, 66). More recently, LKB1, a serine/threonine kinase tumor suppressor (7), was also found to interact with and phosphorylate PTEN in vitro (36). In aggregate, these data suggest that PTEN functional output is controlled by a complex interplay of protein interactions and regulation of C-terminal phosphorylation.Beyond these interactions, there is also evidence to support additional regulatory mechanisms by which the tumor suppressor function of PTEN is mediated. The herpesvirus-associated ubiquitin-specific protease was shown to interact directly with PTEN and promote its nuclear entry (53). Both ubiquitination and relocalization into the nucleus constitute important PTEN regulatory mechanisms (53, 64). In many tumors, PTEN nuclear exclusion has been associated with poor cancer prognosis and more aggressive cancer development (15, 44, 56). Moreover, successful treatment of acute promyelocytic leukemia was shown to be associated with an increase in monoubiquitinylation and relocation of PTEN into the nucleus (53).Like PTEN, the p85 regulatory subunit of PI3K serves as a prominent modulator of PI3K/AKT signaling. p85, which exists in three isoforms (α, β, and γ), targets the catalytic (110-kDa) PI3K subunit to the membrane, which brings it into proximity with membrane-associated phosphatidylinositol lipids. In the steady state, p85 forms a tight association with the catalytic PI3K subunit, usually p110α or p110β in nonhematopoietic cells, with p110δ predominating in leukocytes (19). Consistent with this notion, p85 and p110 exist in equimolar ratios in a wide variety of mammalian cell lines and tissues (19), although some studies have suggested a role for free p85 in cell signaling (33, 65).Several recent lines of evidence have begun to support a possible regulatory relationship between PTEN and p85 (reviewed in references 3 and 53). For example, liver-specific deletion of PIK3R1, which encodes the p85α regulatory subunit, reduces both the activation of PI3K and PTEN enzymatic activity in this context. As a result, p85α-deficient hepatic cells express elevated levels of phosphoinositide trisphosphate and exhibit prolonged AKT activation (60). In addition, both PTEN and p85 are regulated by small GTPase proteins such as RhoA, but PTEN coimmunoprecipitates with the RhoA effector Rock only in the presence of PI3K (18, 31, 37). Although only correlative in nature, these findings may suggest a possible role for PTEN in p85 regulation or vice versa, in addition to its known function as a direct antagonist of the PI3K/AKT pathway (3, 9, 52, 57, 60).In the present study, we demonstrate an endogenous association between p85 and PTEN. Using newly generated antibodies that selectively recognize the PTEN C-terminal tail in its unphosphorylated form, we demonstrate that this PTEN-p85 association preferentially involves the unphosphorylated form of PTEN. The specificity of this interaction was confirmed using multiple antibodies and through studies of both human cancer cells and murine embryonic fibroblasts (MEFs) deficient for specific p85 subunits. This association, which also engages p110β, is enhanced by trastuzumab treatment and correlates with diminished AKT phosphorylation. These results support a functional role for the PTEN-p85 association that may have important biological and therapeutic implications for PI3K/AKT pathway regulation.  相似文献   

4.
5.
6.
The first morphological evidence of African swine fever virus (ASFV) assembly is the appearance of precursor viral membranes, thought to derive from the endoplasmic reticulum, within the assembly sites. We have shown previously that protein p54, a viral structural integral membrane protein, is essential for the generation of the viral precursor membranes. In this report, we study the role of protein p17, an abundant transmembrane protein localized at the viral internal envelope, in these processes. Using an inducible virus for this protein, we show that p17 is essential for virus viability and that its repression blocks the proteolytic processing of polyproteins pp220 and pp62. Electron microscopy analyses demonstrate that when the infection occurs under restrictive conditions, viral morphogenesis is blocked at an early stage, immediately posterior to the formation of the viral precursor membranes, indicating that protein p17 is required to allow their progression toward icosahedral particles. Thus, the absence of this protein leads to an accumulation of these precursors and to the delocalization of the major components of the capsid and core shell domains. The study of ultrathin serial sections from cells infected with BA71V or the inducible virus under permissive conditions revealed the presence of large helicoidal structures from which immature particles are produced, suggesting that these helicoidal structures represent a previously undetected viral intermediate.African swine fever virus (ASFV) (61, 72) is the only known DNA-containing arbovirus and the sole member of the Asfarviridae family (24). Infection by this virus of its natural hosts, the wild swine warthogs and bushpigs and the argasid ticks of the genus Ornithodoros, results in a mild disease, often asymptomatic, with low viremia titers, that in many cases develops into a persistent infection (3, 43, 71). In contrast, infection of domestic pigs leads to a lethal hemorrhagic fever for which the only available methods of disease control are the quarantine of the affected area and the elimination of the infected animals (51).The ASFV genome is a lineal molecule of double-stranded DNA of 170 to 190 kbp in length with convalently closed ends and terminal inverted repeats. The genome encodes more than 150 open reading frames, half of which lack any known or predictable function (16, 75).The virus particle, with an overall icosahedral shape and an average diameter of 200 nm (11), is organized in several concentric layers (6, 11, 15) containing more than 50 structural proteins (29). Intracellular particles are formed by an inner viral core, which contains the central nucleoid surrounded by a thick protein coat, referred to as core shell. This core is enwrapped by an inner lipid envelope (7, 34) on top of which the icosahedral capsid is assembled (26, 27, 31). Extracellular virions possess an additional membrane acquired during the budding from the plasma membrane (11). Both forms of the virus, intracellular and extracellular, are infective (8).The assembly of ASFV particles occurs in the cytoplasm of the infected cell, in viral factories located close to the cell nucleus (6, 13, 49). ASFV factories possess several characteristics similar to those of the cellular aggresomes (35), which are accumulations of aggregates of cellular proteins that form perinuclear inclusions (44).Current models propose that ASFV assembly begins with the modification of endoplasmic reticulum (ER) membranes, which are subsequently recruited to the viral factories and transformed into viral precursor membranes. These ER-derived viral membranes represent the precursors of the inner viral envelope and are the first morphological evidence of viral assembly (7, 60). ASFV viral membrane precursors evolve into icosahedral intermediates and icosahedral particles by the progressive assembly of the outer capsid layer at the convex face of the precursor membranes (5, 26, 27, 31) through an ATP- and calcium-dependent process (19). At the same time, the core shell is formed underneath the concave face of the viral envelope, and the viral DNA and nucleoproteins are packaged and condensed to form the innermost electron-dense nucleoid (6, 9, 12, 69). However, the assembly of the capsid and the internal envelope appears to be largely independent of the components of the core of the particle, since the absence of the viral polyprotein pp220 during assembly produces empty virus-like particles that do not contain the core (9).Comparative genome analysis suggests that ASFV shares a common origin with the members of the proposed nucleocytoplasmic large DNA viruses (NCLDVs) (40, 41). The reconstructed phylogeny of NCLDVs as well as the similitude in the structures and organizations of the genomes indicates that ASFV is more closely related to poxviruses than to other members of the NCLDVs. A consensus about the origin and nature of the envelope of the immature form of vaccinia virus (VV), the prototypical poxvirus, seems to be emerging (10, 17, 20, 54). VV assembly starts with the appearance of crescent-shaped structures within specialized regions of the cytoplasm also known as viral factories (21, 23). The crescent membranes originate from preexisting membranes derived from some specialized compartment of the ER (32, 37, 52, 53, 67), and an operative pathway from the ER to the crescent membrane has recently been described (38, 39). VV crescents apparently grow in length while maintaining the same curvature until they become closed circles, spheres in three dimensions, called immature virions (IV) (22). The uniform curvature is produced by a honeycomb lattice of protein D13L (36, 70), which attaches rapidly to the membranes so that nascent viral membranes always appear to be coated over their entirety. The D13L protein is evolutionarily related to the capsid proteins of the other members of the NCLDV group, including ASFV, but lacks the C-terminal jelly roll motif (40). This structural difference is probably related to the fact that poxviruses are the only member of this group without an icosahedral capsid; instead, the spherical D13L coat acts as a scaffold during the IV stage but is discarded in subsequent steps of morphogenesis (10, 28, 46, 66). Thus, although crescents in VV and precursors of the inner envelope in ASFV are the first morphogenetic stages discernible in the viral factories of these viruses, they seem to be different in nature. Crescents are covered by the D13L protein and are more akin to the icosahedral intermediates of ASFV assembly, whereas ASFV viral membrane precursors are more similar to the naked membranes seen when VV morphogenesis is arrested by rifampin treatment (33, 47, 48, 50) or when the expression of the D13L and A17L proteins are repressed during infection with lethal conditional VV viruses (45, 55, 56, 68, 74, 76).Although available evidence strongly supports the reticular origin of the ASFV inner envelope (7, 60), the mechanism of acquisition remains unknown, and the number of membranes present in the inner envelope is controversial. The traditional view of the inner envelope as formed by two tightly opposed membranes derived from ER collapsed cisternae (7, 59, 60) has recently been challenged by the careful examination of the width of the internal membrane of viral particles and the single outer mitochondrial membrane, carried out using chemical fixation, cryosectioning, and high-pressure freezing (34). The results suggest that the inner envelope of ASFV is a single lipid bilayer, which raises the question of how such a structure can be generated and stabilized in the precursors of the ASFV internal envelope. In the case of VV, the coat of the D13L protein has been suggested to play a key role in the stabilization of the single membrane structure of the crescent (10, 17, 36), but the ASFV capsid protein p72 is not a component of the viral membrane precursors. The identification and functional characterization of the proteins involved in the generation of these structures are essential for the understanding of the mechanisms involved in these early stages of viral assembly. For this reason, we are focusing our interest on the study of abundant structural membrane proteins that reside at the inner envelope of the viral particle. We have shown previously that one of these proteins, p54, is essential for the recruitment of ER membranes to the viral factory (59). Repression of protein p54 expression has a profound impact on virus production and leads to an early arrest in virion morphogenesis, resulting in the virtual absence of membranes in the viral factory.Protein p17, encoded by the late gene D117L in the BA71V strain, is an abundant structural protein (60, 65). Its sequence, which is highly conserved among ASFV isolates (16), does not show any significant similarity with the sequences present in the databases. Protein p17 is an integral membrane protein (18) that is predicted to insert in membranes with a Singer type I topology and has been localized in the envelope precursors as well as in both intracellular and extracellular mature particles (60), suggesting that it resides at the internal envelope, the only membranous structure of the intracellular particles.In this work, we analyze the role of protein p17 in viral assembly by means of an IPTG (isopropyl-β-d-thiogalactopyranoside)-dependent lethal conditional virus. The data presented indicate that protein p17 is essential for viral morphogenesis. The repression of this protein appears to block assembly at the level of viral precursor membranes, resulting in their accumulation at the viral factory.From the electron microscopy analysis of serial sections of viral factories at very early times during morphogenesis, we present experimental evidence that suggests that, during assembly, viral precursor membranes and core material organize into large helicoidal intermediates from which icosahedral particles emerge. The possible role of these structures during ASFV morphogenesis is discussed.  相似文献   

7.
Cell migration is critical for normal development and for pathological processes including cancer cell metastasis. Dynamic remodeling of focal adhesions and the actin cytoskeleton are crucial determinants of cell motility. The Rho family and the mitogen-activated protein kinase (MAPK) module consisting of MEK-extracellular signal-regulated kinase (ERK) are important regulators of these processes, but mechanisms for the integration of these signals during spreading and motility are incompletely understood. Here we show that ERK activity is required for fibronectin-stimulated Rho-GTP loading, Rho-kinase function, and the maturation of focal adhesions in spreading cells. We identify p190A RhoGAP as a major target for ERK signaling in adhesion assembly and identify roles for ERK phosphorylation of the C terminus in p190A localization and activity. These observations reveal a novel role for ERK signaling in adhesion assembly in addition to its established role in adhesion disassembly.Cell migration is a highly coordinated process essential for physiological and pathological processes (69). Signaling through Rho family GTPases (e.g., Rac, Cdc42, and Rho) is crucial for cell migration. Activated Rac and Cdc42 are involved in the production of a dominant lamellipodium and filopodia, respectively, whereas Rho-stimulated contractile forces are required for tail retraction and to maintain adhesion to the matrix (57, 58, 68). Rac- and Cdc42-dependent membrane protrusions are driven by the actin cytoskeleton and the formation of peripheral focal complexes; Rho activation stabilizes protrusions by stimulating the formation of mature focal adhesions and stress fibers. Active Rho influences cytoskeletal dynamics through effectors including the Rho kinases (ROCKs) (2, 3).Rho activity is stimulated by GEFs that promote GTP binding and attenuated by GTPase-activating proteins (GAPs) that enhance Rho''s intrinsic GTPase activity. However, due to the large number of RhoGEFs and RhoGAPs expressed in mammalian cells, the molecular mechanisms responsible for regulation of Rho activity in time and space are incompletely understood. p190A RhoGAP (hereafter p190A) is implicated in adhesion and migration signaling. p190A contains an N-terminal GTPase domain, a large middle domain juxtaposed to the C-terminal GAP domain, and a short C-terminal tail (74). The C-terminal tail of ∼50 amino acids is divergent between p190A and the closely related family member p190B (14) and thus may specify the unique functional roles for p190A and p190B revealed in gene knockout studies (10, 11, 41, 77, 78). p190A activity is dynamically regulated in response to external cues during cell adhesion and migration (5, 6, 59). Arthur et al. (5) reported that p190A activity is required for the transient decrease in RhoGTP levels seen in fibroblasts adhering to fibronectin. p190A activity is positively regulated by tyrosine phosphorylation (4, 5, 8, 17, 31, 39, 40, 42): phosphorylation at Y1105 promotes its association with p120RasGAP and subsequent recruitment to membranes or cytoskeleton (8, 17, 27, 31, 71, 75, 84). However, Y1105 phosphorylation is alone insufficient to activate p190A GAP activity (39). While the functions of p190A can be irreversibly terminated by ubiquitinylation in a cell-cycle-dependent manner (80), less is known about reversible mechanisms that negatively regulate p190A GAP activity during adhesion and motility.The integration of Rho family GTPase and extracellular signal-regulated kinase (ERK) signaling is important for cell motility (48, 50, 63, 76, 79). Several studies have demonstrated a requirement for ERK signaling in the disassembly of focal adhesions in migrating cells, in part through the activation of calpain proteases (36, 37) that can downregulate focal adhesion kinase (FAK) signaling (15), locally suppress Rho activity (52), and sever cytoskeletal linkers to focal adhesions (7, 33). Inhibition of ERK signaling increases focal adhesion size and retards disassembly of focal adhesions in adherent cells (57, 64, 85, 86). It is also recognized that ERK modulates Rho-dependent cellular processes, including membrane protrusion and migration (18, 25, 64, 86). Interestingly, ERK activated in response to acute fibronectin stimulation localizes not only to mature focal adhesions, but also to peripheral focal complexes (32, 76). Since these complexes can either mature or be turned over (12), ERK may play a distinct role in focal adhesion assembly. ERK is proposed to promote focal adhesion formation by activating myosin light chain kinase (MLCK) (21, 32, 50).Here we find that ERK activity is required for Rho activation and focal adhesion formation during adhesion to fibronectin and that p190A is an essential target of ERK signaling in this context. Inspection of the p190A C terminus reveals a number of consensus ERK sites and indeed p190A is phosphorylated by recombinant ERK only on its C terminus in vitro, and on the same C-terminal peptide in vivo. Mutation of the C-terminal ERK phosphorylation sites to alanine increases the biochemical and biological activity of p190A. Finally, inhibition of MEK or mutation of the C-terminal phosphorylation sites enhances retention of p190A in peripheral membranes during spreading on fibronectin. Our data support the conclusion that ERK phosphorylation inhibits p190A allowing increases in RhoGTP and cytoskeletal changes necessary for focal adhesion formation.  相似文献   

8.
9.
To understand how DEXD/H-box proteins recognize and interact with their cellular substrates, we have been studying Prp28p, a DEXD/H-box splicing factor required for switching the U1 snRNP with the U6 snRNP at the precursor mRNA (pre-mRNA) 5′ splice site. We previously demonstrated that the requirement for Prp28p can be eliminated by mutations that alter either the U1 snRNA or the U1C protein, suggesting that both are targets of Prp28p. Inspired by this finding, we designed a bypass genetic screen to specifically search for additional, novel targets of Prp28p. The screen identified Prp42p, Snu71p, and Cbp80p, all known components of commitment complexes, as well as Ynl187p, a protein of uncertain function. To examine the role of Ynl187p in splicing, we carried out extensive genetic and biochemical analysis, including chromatin immunoprecipitation. Our data suggest that Ynl187p acts in concert with U1C and Cbp80p to help stabilize the U1 snRNP-5′ splice site interaction. These findings are discussed in the context of DEXD/H-box proteins and their role in vivo as well as the potential need for more integral U1-snRNP proteins in governing the fungal 5′ splice site RNA-RNA interaction compared to the number of U1 snRNP proteins needed by metazoans.Nuclear precursor mRNA (pre-mRNA) splicing takes place in the spliceosome, a large dynamic complex consisting of over 100 proteins and five small nuclear RNAs (snRNAs) (32, 70). During spliceosome assembly, the U1 small nuclear ribonucleoprotein particle (snRNP) first contacts the pre-mRNA 5′ splice site (5′ss), followed by binding of the U2 snRNP to the branch site and the joining of the U5-U4/U6 tri-snRNP (32, 64, 70). The step in which U1 snRNP binds to the 5′ss is arguably one of the most critical, because it probably commits pre-mRNA to the splicing pathway (38, 48, 49, 60, 74). In the budding yeast Saccharomyces cerevisiae in vitro system, two U1-snRNP-containing commitment complexes (CCs), CC1 and CC2, can be detected by native gel electrophoresis prior to the U2 snRNP''s joining to form the prespliceosome (38, 60). CC1, whose formation is dependent on a functional 5′ss, appears to be a kinetic precursor to CC2, whose formation requires both a functional 5′ss and branch site and the participation of the branch-site-binding protein (BBP) and Mud2p, which are likely equivalent to SF1 and U2AF65, respectively, in the mammalian system (1-3, 75).Accumulating evidence suggests that formation of the canonical 5- to 7-bp RNA duplex between U1 snRNA and the 5′ss region is not sufficient to cause a stable CC to form in the yeast system (59, 62, 78); protein-RNA contacts are also important. For example, Zhang and Rosbash (77) identified eight proteins, all present in CCs, that make physical contact with the pre-mRNA at or near the 5′ss. Four of these proteins, U1C, U1-70K, Snu56p, and Nam8p, are integral parts of the U1 snRNP (20), and another three, SmB, SmD1, and SmD3, belong to the seven-member ring that binds the conserved Sm site present on U1, U2, U4, and U5 snRNAs (33, 71). The remaining protein, Cbp80p, is a subunit of the nuclear cap-binding complex (CBC), which also contains a second subunit, Cbp20p (28, 39). Interestingly, despite being a non-snRNP factor, Cbp80p is known to collaborate with U1 snRNP to help form or stabilize CC1 (8, 40). Furthermore, the contact between the C-terminal tails of SmB, SmD1, and SmD3 and the pre-mRNA may contribute to stabilizing the U1 snRNP/pre-mRNA interaction (76). Finally, Du and Rosbash (11) more recently showed that U1C is capable of selecting splice-site-like sequences in which the first four nucleotides, GUAU, are identical to the first four nucleotides of the yeast splice-site consensus sequence.Once fully assembled, the spliceosome must progress through a number of major structural and conformational changes to form the catalytic center; these include a series of highly orchestrated RNA-RNA rearrangements (53, 64, 70). Some of these are mutually exclusive; i.e., the formation of one RNA duplex requires the disruption of another. For example, the base-pairing interaction between the U1 snRNA and the 5′ss is replaced by a U6 snRNA/5′ss pairing. This exchange appears to be coupled to U4/U6 RNA unwinding (53, 64, 70). It is now known that splicing factors belonging to the ATPase II superfamily (18), which are also termed the DEXD/H-box proteins (5, 43), promote spliceosomal RNA rearrangements (64). However, the precise roles of most DEXD/H-box proteins remain unclear.It has been nearly 2 decades since DEXD/H-box proteins were first proposed to be RNA helicases (44). Over the years, a wealth of data revealed that DEXD/H-box proteins are essential in most, if not all, RNA-related pathways, e.g., splicing, mRNA export, and ribosomal biogenesis (5, 43, 64). Their modes of action in vivo remain a mystery, however. For example, Lorsch and Herschlag (45, 46) proposed that DEXD/H-box proteins may perform functions which are distinct from RNA unwinding and include mediating large-scale RNA structural rearrangements, disrupting protein-RNA or protein-protein interactions, and functioning as fidelity sensors in RNA-RNA interactions and rearrangements. Indeed, recent data confirm that DEXD/H-box proteins can catalyze protein displacement in a manner independent of RNA duplex unwinding (30). Therefore, the essential functions of DExH/D proteins can be exerted on a wide range of RNP substrates. This “RNPase” (or ATPase for RNP remodeling) hypothesis appears especially attractive in light of the fact that RNA duplexes in vivo are rarely more than ∼10 contiguous base pairs in length and that they often require protein binding for stabilization (21, 63). To fully understand how DEXD/H-box proteins function in the cell, it is critical to identify their physiological substrates.Inspired by our previous finding that the requirement for Prp28p, an essential DEXD/H-box splicing factor, can be bypassed by mutations that alter the YHC1 gene, which encodes U1C protein (7), we sought to exploit the bypass concept to deepen our understanding of the role of Prp28p in splicing. The underlying hypothesis is that bypass mutations define gene products that Prp28p may counteract. Here we describe the outcome of this approach and provide a detailed analysis of Ynl187p, a novel protein that probably contributes to stabilizing the U1 snRNP-5′ss interaction.  相似文献   

10.
To facilitate the release of infectious progeny virions, human immunodeficiency virus type 1 (HIV-1) exploits the Endosomal Sorting Complex Required for Transport (ESCRT) pathway by engaging Tsg101 and ALIX through late assembly (L) domains in the C-terminal p6 domain of Gag. However, the L domains in p6 are known to be dispensable for efficient particle production by certain HIV-1 Gag constructs that have the nucleocapsid (NC) domain replaced by a foreign dimerization domain to substitute for the assembly function of NC. We now show that one such L domain-independent HIV-1 Gag construct (termed ZWT) that has NC-p1-p6 replaced by a leucine zipper domain is resistant to dominant-negative inhibitors of the ESCRT pathway that block HIV-1 particle production. However, ZWT became dependent on the presence of an L domain when NC-p1-p6 was restored to its C terminus. Furthermore, when the NC domain was replaced by a leucine zipper, the p1-p6 region, but not p6 alone, conferred sensitivity to inhibition of the ESCRT pathway. In an authentic HIV-1 Gag context, the effect of an inhibitor of the ESCRT pathway on particle production could be alleviated by deleting a portion of the NC domain together with p1. Together, these results indicate that the ESCRT pathway dependence of HIV-1 budding is determined, at least in part, by the NC-p1 region of Gag.Human immunodeficiency virus type 1 (HIV-1) and other retroviruses hijack the cellular Endosomal Sorting Complex Required for Transport (ESCRT) pathway to promote the detachment of virions from the cell surface and from each other (3, 21, 42, 44, 47). The ESCRT pathway was initially identified based on its requirement for the sorting of ubiquitinated cargo into multivesicular bodies (MVB) (50, 51). During MVB biogenesis, the ESCRT pathway drives the membrane deformation and fission events required for the inward vesiculation of the limiting membrane of this organelle (26, 29, 50, 51). More recently, it emerged that the ESCRT pathway is also essential for the normal abscission of daughter cells during the final stage of cell division (10, 43). Most of the components of the ESCRT pathway are involved in the formation of four heteromeric protein complexes termed ESCRT-0, ESCRT-I, ESCRT-II, and ESCRT-III. Additional components include ALIX, which interacts both with ESCRT-I and ESCRT-III, and the AAA ATPase Vps4, which mediates the disassembly of ESCRT-III (29, 42).The deformation and scission of endocytic membranes is thought to be mediated by ESCRT-III, which, together with Vps4, constitutes the most conserved element of the pathway (23, 26, 42). Indeed, it was recently shown that purified yeast ESCRT-III induces membrane deformation (52), and in another study three subunits of yeast ESCRT-III were sufficient to promote the formation of intralumenal vesicles in an in vitro assay (61). In mammals, ESCRT-III is formed by the charged MVB proteins (CHMPs), which are structurally related and tightly regulated through autoinhibition (2, 33, 46, 53, 62). The removal of an inhibitory C-terminal domain induces polymerization and association with endosomal membranes and converts CHMPs into potent inhibitors of retroviral budding (34, 46, 53, 60, 62). Alternatively, CHMPs can be converted into strong inhibitors of the ESCRT pathway and of HIV-1 budding through the addition of a bulky tag such as green fluorescent protein (GFP) or red fluorescent protein (RFP) (27, 36, 39, 54). Retroviral budding in general is also strongly inhibited by catalytically inactive Vps4 (22, 41, 55), or upon Vsp4B depletion (31), confirming the crucial role of ESCRT-III.Retroviruses engage the ESCRT pathway through the activity of so-called late assembly (L) domains in Gag. In the case of HIV-1, the primary L domain maps to a conserved PTAP motif in the C-terminal p6 domain of Gag (24, 28) and interacts with the ESCRT-I component Tsg101 (15, 22, 40, 58). HIV-1 p6 also harbors an auxiliary L domain of the LYPxnL type, which interacts with the V domain of ALIX (20, 35, 39, 54, 59, 63). Interestingly, Tsg101 binding site mutants of HIV-1 can be fully rescued through the overexpression of ALIX, and this rescue depends on the ALIX binding site in p6 (20, 56). In contrast, the overexpression of a specific splice variant of the ubiquitin ligase Nedd4-2 has been shown to rescue the release and infectivity of HIV-1 mutants lacking all known L domains in p6 (12, 57). Nedd4 family ubiquitin ligases had previously been implicated in the function of PPxY-type L domains, which also depend on an intact ESCRT pathway for function (4, 32, 38). However, HIV-1 Gag lacks PPxY motifs, and the WW domains of Nedd4-2, which mediate its interaction with PPxY motifs, are dispensable for the rescue of HIV-1 L domain mutants (57).ALIX also interacts with the nucleocapsid (NC) region of HIV-1 Gag (18, 49), which is located upstream of p6 and the p1 spacer peptide. ALIX binds HIV-1 NC via its Bro1 domain, and the capacity to interact with NC and to stimulate the release of a minimal HIV-1 Gag construct is shared among widely divergent Bro1 domain proteins (48). Based on these findings and the observation that certain mutations in NC cause a phenotype that resembles that of L domain mutants, it has been proposed that NC cooperates with p6 to recruit the machinery required for normal HIV-1 budding (18, 49).NC also plays a role in Gag polyprotein multimerization, and this function of NC depends on its RNA-binding activity (5-8). It has been proposed that the role of the NC-nucleic acid interaction during assembly is to promote the formation of Gag dimers (37), and HIV-1 assembly in the absence of NC can indeed be efficiently rescued by leucine zipper dimerization domains (65). Surprisingly, in this setting the L domains in p6 also became dispensable, since particle production remained efficient even when the entire NC-p1-p6 region of HIV-1 Gag was replaced by a leucine zipper (1, 65). These findings raised the possibility that the reliance of wild-type (WT) HIV-1 Gag on a functional ESCRT pathway is, at least in part, specified by NC-p1-p6. However, it also remained possible that the chimeric Gag constructs engaged the ESCRT pathway in an alternative manner.In the present report, we provide evidence supporting the first of those two possibilities. Particle production became independent of ESCRT when the entire NC-p1-p6 region was replaced by a leucine zipper, and reversion to ESCRT dependence was shown to occur as a result of restoration of p1-p6 but not of p6 alone. Furthermore, although the deletion of p1 alone had little effect in an authentic HIV-1 Gag context, the additional removal of a portion of NC improved particle production in the presence of an inhibitor of the ESCRT pathway. Together, these data imply that the NC-p1 region plays an important role in the ESCRT-dependence of HIV-1 particle production.  相似文献   

11.
12.
13.
14.
15.
p32/gC1qR/C1QBP/HABP1 is a mitochondrial/cell surface protein overexpressed in certain cancer cells. Here we show that knocking down p32 expression in human cancer cells strongly shifts their metabolism from oxidative phosphorylation (OXPHOS) to glycolysis. The p32 knockdown cells exhibited reduced synthesis of the mitochondrial-DNA-encoded OXPHOS polypeptides and were less tumorigenic in vivo. Expression of exogenous p32 in the knockdown cells restored the wild-type cellular phenotype and tumorigenicity. Increased glucose consumption and lactate production, known as the Warburg effect, are almost universal hallmarks of solid tumors and are thought to favor tumor growth. However, here we show that a protein regularly overexpressed in some cancers is capable of promoting OXPHOS. Our results indicate that high levels of glycolysis, in the absence of adequate OXPHOS, may not be as beneficial for tumor growth as generally thought and suggest that tumor cells use p32 to regulate the balance between OXPHOS and glycolysis.Tumors can be distinguished from their nonmalignant counterparts by specific molecular signatures expressed in malignant cells and tumor vasculature. We explore such differences by identifying tumor-homing peptides from phage libraries that we screen in vivo (60). We recently showed (19) that the cellular receptor for one of our tumor-homing peptides is a protein variously known as p32, p33, gC1q receptor (gC1qR), or hyaluronic acid binding protein 1 (HABP1). This protein was originally isolated based on its copurification with the nuclear splicing factor SF-2 (37). However, it was subsequently shown to bind also the globular heads of complement component C1q (23), hyaluronic acid (10), and numerous other extracellular and intracellular proteins (24, 28, 33, 42). Most recently it has been shown that p32 interacts with the long and short forms of the tumor suppressor ARF (30, 56, 57). Despite the numerous reports on p32 interaction partners, the role of these binding activities in the physiological function of the protein is unknown, and some investigators have proposed that p32 may be a chaperone protein (58, 65).The p32 protein is primarily localized in the mitochondrial matrix (12, 46, 48) but has also been reported to be present in other subcellular locations (53). Some of the p32 protein can be at the cell surface, a location that appears to be specific for tumors (19). In this regard, p32 is similar to some other intracellular proteins that are also partially localized at the cell surface in tumor cells (8, 49). In addition to the partial cell surface localization of p32, many human tumors exhibit higher p32 expression levels than their nonmalignant counterpart tissues (7, 19, 52, 59). Moreover, p32 is differentially expressed during the progression of epidermal carcinoma, accumulating in metastatic islands (25).We set out to modulate p32 expression in tumor cells to gain information on the role of this protein in cancer. We show here that p32 knockdown cells shift their metabolism from oxidative phosphorylation (OXPHOS) toward glycolysis and become poorly tumorigenic. These changes could be reversed by restored expression of p32. These results show that p32 supports oxidative phosphorylation in human cancer cells and opposes the shift of tumor cell metabolism toward glycolysis. The unique expression pattern of p32 in tumors and its crucial role in tumor metabolism make p32 a promising target for tumor therapy. The fact that this protein is upregulated in tumors and counteracts glycolytic metabolism suggests that the role of the Warburg effect in tumor growth may not be as straightforward as is generally thought.  相似文献   

16.
17.
Plant viral infection and spread depends on the successful introduction of a virus into a cell of a compatible host, followed by replication and cell-to-cell transport. The movement proteins (MPs) p8 and p9 of Turnip crinkle virus are required for cell-to-cell movement of the virus. We have examined the membrane association of p9 and found that it is an integral membrane protein with a defined topology in the endoplasmic reticulum (ER) membrane. Furthermore, we have used a site-specific photo-cross-linking strategy to study the membrane integration of the protein at the initial stages of its biosynthetic process. This process is cotranslational and proceeds through the signal recognition particle and the translocon complex.Cell-to-cell transport of plant virus requires the virally encoded movement proteins (MPs). These proteins specialize in the translocation of the viral genome or, in some cases, the virions from the replication/encapsidation site to adjacent cells. This process takes place through the plasmodesmata (PD), the small pores formed by prolongations of the endoplasmic reticulum (ER) membranes trapped within the center of the plasma membrane-lined cytoplasmic cylinder that connect plant cells. MPs belong to different protein families with unique functional and structural characteristics. The most studied MP is p30 from the Tobacco mosaic virus, a 30-kDa RNA-binding protein (4) with two putative transmembrane (TM) segments (2) that has so far been considered an integral membrane protein (13, 42). At an early stage of infection, p30 associates with the ER network (18, 59). Given that the ER is continuous through PD, it was suggested that the movement complex transports cell to cell via the PD. On the other hand, passage through the connecting structure largely remains a mystery, although it seems reasonable that the process again occurs in close juxtaposition to the ER-derived membrane (desmotubule) that runs through the PD (12, 35). Many other plant viruses have a cell-to-cell transport system based not on one but on two (double-gene block [DGB]) or even three (triple-gene block [TGB]) MPs. In some of these cases it has been shown that at least one MP is closely associated with the ER membrane (28, 34, 41, 50, 55). Thus, it has been assumed that other MPs associate similarly with membranes.The targeting and insertion of an integral membrane protein can occur either posttranslationally, in which the protein is completely synthesized on cytosolic ribosomes before being inserted, or cotranslationally, in which protein synthesis and integration into the ER membrane are coupled. In the latter case, the targeting of the ribosome-mRNA-nascent chain complex to the membrane depends on the signal recognition particle (SRP) and its interaction with the membrane-bound SRP receptor (11), which is located in close proximity to the translocon. The translocon, a multiprotein complex composed of the Sec61α, -β, and -γ subunits (16) and the translocating chain-associated membrane protein (TRAM) (15) in eukaryotic cells, facilitates the translocation of soluble proteins into the ER lumen and the insertion of integral membrane proteins into the lipid bilayer (24).Plant virus infection depends on the proper targeting and association or insertion of the movement proteins with or into the ER membrane. In this report, we investigate the insertion into, topology of, and targeting to the membrane of the p9 MP from Turnip crinkle virus (TCV). This is a positive-sense single-stranded RNA virus that belongs to the Carmovirus genus and thus to the DGB. Its 4-kb genome encodes five open reading frames (ORFs) (3, 17). Translation of the first two yields p28 and p88, both implicated in viral RNA synthesis. In the central region, two overlapping ORFs encode the small proteins p8 and p9, which have been shown to be involved in cell-to-cell movement (6, 17, 31). The RNA-binding protein p8 (17, 58) overlaps the distal 3′ region of the replicase p88. The 3′ region of the genome encodes the viral coat protein p38, and its 5′ end overlaps p9 (3).A strong interaction with the membrane is expected for p9 due to the close similarities in the genomic arrangement of TCV (57) with other carmoviruses, like Carnation mottle virus (CarMV) and Melon necrotic spot virus (MNSV). Both CarMV and MNSV have two small MPs, one an RNA-binding protein (39, 53, 54) and the other a cotranslationally inserted integral membrane protein (34, 47, 55). In this study, we present evidence of the integration of TCV p9 into ER-derived microsomal membranes. Using an in vitro translation system based on a model integral membrane protein, we have been able to identify two membrane-spanning domains. Additionally, the membrane topology of the p9 MP was analyzed in vitro and found to have an N terminus (N-t)/C terminus (C-t) luminal orientation. Finally, using a site-directed photo-cross-linking approach, we demonstrated that the mechanism of p9 insertion into the ER membrane involves SRP and the translocon.  相似文献   

18.
Human immunodeficiency virus type 1 (HIV-1) Nef interferes with the endocytic machinery to modulate the cell surface expression of CD4. However, the basal trafficking of CD4 is governed by different rules in the target cells of HIV-1: whereas CD4 is rapidly internalized from the cell surface in myeloid cells, CD4 is stabilized at the plasma membrane through its interaction with the p56lck kinase in lymphoid cells. In this study, we showed that Nef was able to downregulate CD4 in both lymphoid and myeloid cell lines but that an increase in the internalization rate of CD4 could be observed only in lymphoid cells. Expression of p56lck in nonlymphoid CD4-expressing cells restores the ability of Nef in order to increase the internalization rate of CD4. Concurrent with this observation, the expression of a p56lck-binding-deficient mutant of CD4 in lymphoid cells abrogates the Nef-induced acceleration of CD4 internalization. We also show that the expression of Nef causes a decrease in the association of p56lck with cell surface-expressed CD4. Regardless of the presence of p56lck, the downregulation of CD4 by Nef was followed by CD4 degradation. Our results imply that Nef uses distinct mechanisms to downregulate the cell surface expression levels of CD4 in either lymphoid or myeloid target cells of HIV-1.Besides proteins that are essential for proper virus processing and assembly, the genomes of primate lentiviruses such as human immunodeficiency virus type 1 (HIV-1) encode auxiliary proteins that modulate viral infectivity. The 27-kDa auxiliary protein Nef is a key element in the progression of primary HIV-1 infection toward AIDS. Cases of patients infected with HIV-1 strains harboring a deletion in the nef gene or a defective nef allele have been reported. Some of these patients exhibit asymptomatic or slow progression toward the disease (6, 17, 37). In vitro, Nef facilitates viral replication and enhances the infectivity of viral particles (13, 47, 69). The mechanisms involved in the Nef-induced increase of viral infectivity remain elusive; however, it is a multifactorial process related to the ability of Nef to alter the trafficking of host cell proteins.Indeed, the most documented effect of Nef during the course of viral infection is its ability to disturb the clathrin-dependent trafficking machinery involved in the transport of transmembrane proteins through endosomal compartments. This leads to the modulation of the level of cell surface expression for some receptors, including CD4, which is the primary receptor of HIV-1 (35) and major histocompatibility complex class I (reviewed in references 22 and 27). The downregulation of CD4, which results in the impairment of the immunological synapse (72) and the downregulation of major histocompatibility complex class I molecules (reviewed in reference 16), is believed to contribute to the escape of HIV-1-infected cells from immunosurveillance. Moreover, the downregulation of CD4 helps avoid superinfection of cells, which would be deleterious to the virus (reviewed in reference 21), and has a direct impact on viral fitness by allowing better incorporation of the functional envelope in viral particles produced from CD4-expressing cells (3, 36, 53).Nef-induced cell surface downregulation of CD4 is efficient in all CD4-expressing cells and depends on the integrity of a di-Leu motif at position 164/165 of the C-terminal flexible loop of HIV-1 Nef (2, 9, 25). This di-Leu motif allows for the interaction with clathrin-associated adaptor protein (AP) complexes that participate in the clathrin-dependent vesicular transport within the endocytic pathway. The AP type 2 (AP-2) complex is localized at the plasma membrane and is essential to the assembly and function of clathrin-coated pits involved in the internalization of receptors from the cell surface (59). The interaction of Nef with AP-2 is well delineated and has been proposed to enhance the targeting of CD4 to clathrin-coated pits and its internalization (10, 12, 26, 32, 39).Helper T lymphocytes are the predominant cell type that expresses CD4; however, CD4 is also present at the surfaces of monocytes and macrophages (70), where its function is yet to be elucidated. Whereas cell surface CD4 is rapidly internalized in myeloid cells, CD4 is stabilized at the plasma membrane in lymphoid cells through its interaction with the Src family protein tyrosine kinase p56lck. Cys residues located at positions 420/422 in the CD4 cytoplasmic tail are essential to the constitutive association with p56lck (73). Besides its role in signal transduction, this interaction also correlates with an accumulation of CD4 in lipid rafts and enhanced exclusion of CD4 from clathrin-coated pits (50).In T cells, treatment with phorbol esters such as phorbol 12-myristate 13-acetate (PMA) provokes the phosphorylation of Ser residues found in the cytoplasmic tail of CD4. This correlates with a decreased association of p56lck with CD4 and the internalization of the receptor (24, 32-34, 41, 45, 48, 52, 56, 61, 66-68). Nef-induced CD4 downregulation is known to be independent of Ser phosphorylation (20) and is therefore governed by mechanisms different from those involved in PMA-induced CD4 downregulation. However, the Leu-based sorting motif in the CD4 cytoplasmic tail is critical for both PMA and Nef-induced CD4 downregulation (2, 5, 24, 31, 56, 60, 68), thus indicating that despite being different, the mechanisms involved in Nef- and PMA-induced CD4 downregulation partially overlap.In the present study, we investigated whether the mechanisms used by Nef to downregulate CD4 are cell type-dependent processes. We looked at the trafficking and steady-state expression of CD4 in the main target cells of HIV-1, CD4-positive T lymphocytes, and cells of the monocyte/macrophage lineage. Our results demonstrate that the presence of p56lck has a direct impact on the mechanisms used by Nef to downregulate CD4 from the cell surface of T lymphocytes. They also reveal that Nef uses distinct pathways to decrease levels of cell surface expression of CD4 in lymphoid or myeloid target cells of HIV-1.  相似文献   

19.
20.
In many human cancers, p27 downregulation correlates with a worse prognosis, suggesting that p27 levels could represent an important determinant in cell transformation and cancer development. Using a mouse model system based on v-src-induced transformation, we show here that p27 absence is always linked to a more aggressive phenotype. When cultured in three-dimensional contexts, v-src-transformed p27-null fibroblasts undergo a morphological switch from an elongated to a rounded cell shape, accompanied by amoeboid-like morphology and motility. Importantly, the acquisition of the amoeboid motility is associated with a greater ability to move and colonize distant sites in vivo. The reintroduction of different p27 mutants in v-src-transformed p27-null cells demonstrates that the control of cell proliferation and motility represents two distinct functions of p27, both necessary for it to fully act as a tumor suppressor. Thus, we highlight here a new p27 function in driving cell plasticity that is associated with its C-terminal portion and does not depend on the control of cyclin-dependent kinase activity.Dissemination of tumor cells is strictly linked to their ability to attach to and move within the extracellular matrix (ECM) in a three-dimensional (3D) environment. The use of 3D experimental model systems revealed that a higher complexity in cell migration and adaptation responses exists in the 3D model than in the classical 2D model (10, 16, 41, 49). A striking example is given by the fact that only in 3D could individually migrating cells use different mechanisms such as mesenchymal and amoeboid motility (16, 17). The relative slow mesenchymal migration is characterized by a fibroblast-like spindle shape and is dependent on integrin-mediated adhesion and on protease function (16). The amoeboid motility can in some cases represent a less adhesive, integrin-independent type of movement. Cells use a propulsive mechanism and are highly deformable, and rather than degrade the matrix, they are able to squeeze through it (16). As a result, the cells that use the amoeboid motility can potentially move faster than cells that use a mesenchymal strategy. Mesenchymal and amoeboid movements are also characterized by a different involvement of small GTPases of the Rho family. A high RhoA activity is associated mainly with the amoeboid motility, while the mesenchymal migration needs a high Rac activity at the leading edge to promote the extension of cellular protrusions (41, 48). Under certain circumstances, cancer cells can undergo conversion from a mesenchymal toward an amoeboid motility, an event referred as mesenchymal-amoeboid transition (MAT) (50). MAT represents a putative escape mechanism in tumor cell dissemination that could be induced by inhibition of pericellular proteolysis (50) or by increased membrane-associated RhoA activity (18, 40).Key mediators of cell motility through ECM substrates are the members of the Src family kinases. The prototype of Src family kinases, c-Src (14), is activated following cell-ECM adhesion and contributes to regulate the focal adhesion turnover and the cytoskeletal modifications necessary for normal cell adhesion and motility (52). The c-Src gene is the proto-oncogene of the transforming gene v-src of Rous sarcoma virus, and its elevated protein level and activity have been found in many human tumors (20, 28, 27, 34). Despite the accumulation of information and new molecular understanding of how Src is controlled, there is still an incomplete picture about its role in the generation of the malignant phenotype. v-Src shows higher levels of the kinase activity and transforming ability than c-Src (14, 15, 52). It induces normal cells to acquire a variety of transformed features, including alteration of morphology and increase of invasion ability due to its role in focal adhesion remodeling (7, 9, 13).Many data suggest that there is a close relationship between cell-ECM interaction and the proliferation and movements in both normal and tumor cells (5, 38, 43). Accordingly, Src activation may influence not only cell motility but also cell cycle progression by targeting the cell cycle inhibitor p27kip1 to proteasomal degradation (22, 39). Recent evidences indicated that p27kip1 (hereafter called p27) can also regulate cell migration, even though its role still remains controversial since it has been reported to either block or stimulate cell movements (1, 4, 11, 19, 21, 23, 29, 45).Based on these notions, we tested the possible contribution of p27 to the growth and motility phenotypes induced by v-src transformation, with special regard to those cellular invasive features that can be observed in 3D environments. By studying in vitro and in vivo the behavior of wild-type (WT) and p27-null fibroblasts transformed with v-src, we highlight a new role for p27 in the regulation of cellular plasticity that can ultimately drive tumor cell shape, motility, and invasion.  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