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.
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   

2.
A new role is reported for CP12, a highly unfolded and flexible protein, mainly known for its redox function with A4 glyceraldehyde-3-phosphate dehydrogenase (GAPDH). Both reduced and oxidized CP12 can prevent the in vitro thermal inactivation and aggregation of GAPDH from Chlamydomonas reinhardtii. This mechanism is thus not redox-dependent. The protection is specific to CP12, because other proteins, such as bovine serum albumin, thioredoxin, and a general chaperone, Hsp33, do not fully prevent denaturation of GAPDH. Furthermore, CP12 acts as a specific chaperone, since it does not protect other proteins, such as catalase, alcohol dehydrogenase, or lysozyme. The interaction between CP12 and GAPDH is necessary to prevent the aggregation and inactivation, since the mutant C66S that does not form any complex with GAPDH cannot accomplish this protection. Unlike the C66S mutant, the C23S mutant that lacks the N-terminal bridge is partially able to protect and to slow down the inactivation and aggregation. Tryptic digestion coupled to mass spectrometry confirmed that the S-loop of GAPDH is the interaction site with CP12. Thus, CP12 not only has a redox function but also behaves as a specific “chaperone-like protein” for GAPDH, although a stable and not transitory interaction is observed. This new function of CP12 may explain why it is also present in complexes involving A2B2 GAPDHs that possess a regulatory C-terminal extension (GapB subunit) and therefore do not require CP12 to be redox-regulated.CP12 is a small 8.2-kDa protein present in the chloroplasts of most photosynthetic organisms, including cyanobacteria (1, 2), higher plants (3), the diatom Asterionella formosa (4, 5), and green (1) and red algae (6). It allows the formation of a supramolecular complex between phosphoribulokinase (EC 2.7.1.19) and glyceraldehyde-3-phosphate dehydrogenase (GAPDH),3 two key enzymes of the Calvin cycle pathway, and was recently shown to interact with fructose bisphosphate aldolase, another enzyme of the Calvin cycle pathway (7). The phosphoribulokinase·GAPDH·CP12 complex has been extensively studied in Chlamydomonas reinhardtii (8, 9) and in Arabidopsis thaliana (10, 11). In the green alga C. reinhardtii, the interaction between CP12 and GAPDH is strong (8). GAPDH may exist as a homotetramer composed of four GapA subunits (A4) in higher plants, cyanobacteria, and green and red algae (6, 12), but in higher plants, it can also exist as a heterotetramer (A2B2), composed of two subunits, GapA and GapB (13, 14). GapB, up to now, has exclusively been found in Streptophyta, but recently two prasinophycean green algae, Ostreococcus tauri and Ostreococcus lucimarinus, were also shown to possess a GapB gene, whereas CP12 is missing (15). The GapB subunit is similar to the GapA subunit but has a C-terminal extension containing two redox-regulated cysteine residues (16). Thus, although the A4 GAPDHs lack these regulatory cysteine residues (13, 14, 1720), they are also redox-regulated through its interaction with CP12, since the C terminus of this small protein resembles the C-terminal extension of the GapB subunit. The regulatory cysteine residues for GapA are thus supplied by CP12, as is well documented in the literature (1, 8, 11, 16).CP12 belongs to the family of intrinsically unstructured proteins (IUPs) (2126). The amino acid composition of these proteins causes them to have no or few secondary structures. Their total or partial lack of structure and their high flexibility allow them to be molecular adaptors (27, 28). They are often able to bind to several partners and are involved in most cellular functions (29, 30). Recently, some IUPs have been described in photosynthetic organisms (31, 32).There are many functional categories of IUPs (22, 33). They can be, for instance, involved in permanent binding and have (i) a scavenger role, neutralizing or storing small ligands; (ii) an assembler role by forming complexes; and (iii) an effector role by modulating the activity of a partner molecule (33). These functions are not exclusive; thus, CP12 can form a stable complex with GAPDH, regulating its redox properties (8, 34, 35), and can also bind a metal ion (36, 37). IUPs can also bind transiently to partners, and some of them have been found to possess a chaperone activity (31, 38). This chaperone function was first shown for α-synuclein (39) and for α-casein (40), which are fully disordered. The amino acid composition of IUPs is less hydrophobic than those of soluble proteins; hence, they lack hydrophobic cores and do not become insoluble when heated. Since CP12 belongs to this family, we tested if it was resistant to heat treatment and finally, since it is tightly bound to GAPDH, if it could prevent aggregation of its partner, GAPDH, an enzyme well known for its tendency to aggregate (4144) and consequently a substrate commonly used in chaperone studies (45, 46).Unlike chaperones, which form transient, dynamic complexes with their protein substrates through hydrophobic interactions (47, 48), CP12 forms a stable complex with GAPDH. The interaction involves the C-terminal part of the protein and the presence of negatively charged residues on CP12 (35). However, only a site-directed mutagenesis has been performed to characterize the interaction site on GAPDH. Although the mutation could have an indirect effect, the residue Arg-197 was shown to be a good candidate for the interaction site (49).In this report, we accordingly used proteolysis experiments coupled with mass spectrometry to detect which regions of GAPDH are protected by its association with CP12. To conclude, the aim of this report was to characterize a chaperone function of CP12 that had never been described before and to map the interaction site on GAPDH using an approach that does not involve site-directed mutagenesis.  相似文献   

3.
4.
5.
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   

6.
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   

7.
8.
9.
10.
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   

11.
12.
13.
14.
The 26 S proteasome is a large proteolytic machine, which degrades most intracellular proteins. We found that thioredoxin, Txnl1/TRP32, binds to Rpn11, a subunit of the regulatory complex of the human 26 S proteasome. Txnl1 is abundant, metabolically stable, and widely expressed and is present in the cytoplasm and nucleus. Txnl1 has thioredoxin activity with a redox potential of about-250 mV. Mutant Txnl1 with one active site cysteine replaced by serine formed disulfide bonds to eEF1A1, a substrate-recruiting factor of the 26 S proteasome. eEF1A1 is therefore a likely physiological substrate. In response to knockdown of Txnl1, ubiquitin-protein conjugates were moderately stabilized. Hence, Txnl1 is the first example of a direct connection between protein reduction and proteolysis, two major intracellular protein quality control mechanisms.Degradation of proteins in eukaryotic cells plays a pivotal role in the regulation of several important processes, including cell division, antigen presentation, and signal transduction (1). Most intracellular proteins are degraded by the 26 S proteasome, a 2.5-MDa protease complex composed of more than 30 different subunits (2).To become degraded, proteins are typically first conjugated to a chain of ubiquitin moieties. This reaction is catalyzed by ubiquitin ligases. The ubiquitin chains lend the proteins affinity for the 26 S proteasome (3). For efficient degradation, certain ubiquitylated proteins are shuttled to the 26 S proteasome by substrate recruiting factors, such as Rad23, Dsk2, and eEF1A (4, 5).The 26 S proteasome is composed of two stable subcomplexes, the proteolytically active 20 S core and 19 S regulatory complexes, which bind to one or both ends of the cylindrical 20 S core particle (6). The regulatory complexes first recognize the ubiquitylated substrates (3), before the substrates are deubiquitylated (7, 8), unfolded (9, 10), and translocated into the 20 S particle for degradation.Although the 26 S proteasome has been known for more than 20 years (11), novel subunits and cofactors have been described recently (12, 13). Here we report another novel proteasome-associated protein, Txnl1 (thioredoxin-like protein 1), that associates directly with the proteasome subunit Rpn11. Txnl1 exhibits thioredoxin activity and targets eEF1A1 in vivo. Previous reports have shown that eEF1A1 transfers misfolded nascent proteins from the ribosome to the 26 S proteasome for degradation (5, 14, 15). Accordingly, ubiquitin-protein conjugates were stabilized upon knockdown of Txnl1 expression. Txnl1 therefore directly links protein reduction and proteolysis, two major intracellular protein quality control mechanisms.  相似文献   

15.
The antioxidant N-acetyl-l-cysteine prevented the autophagy-dependent delivery of mitochondria to the vacuoles, as examined by fluorescence microscopy of mitochondria-targeted green fluorescent protein, transmission electron microscopy, and Western blot analysis of mitochondrial proteins. The effect of N-acetyl-l-cysteine was specific to mitochondrial autophagy (mitophagy). Indeed, autophagy-dependent activation of alkaline phosphatase and the presence of hallmarks of non-selective microautophagy were not altered by N-acetyl-l-cysteine. The effect of N-acetyl-l-cysteine was not related to its scavenging properties, but rather to its fueling effect of the glutathione pool. As a matter of fact, the decrease of the glutathione pool induced by chemical or genetical manipulation did stimulate mitophagy but not general autophagy. Conversely, the addition of a cell-permeable form of glutathione inhibited mitophagy. Inhibition of glutathione synthesis had no effect in the strain Δuth1, which is deficient in selective mitochondrial degradation. These data show that mitophagy can be regulated independently of general autophagy, and that its implementation may depend on the cellular redox status.Autophagy is a major pathway for the lysosomal/vacuolar delivery of long-lived proteins and organelles, where they are degraded and recycled. Autophagy plays a crucial role in differentiation and cellular response to stress and is conserved in eukaryotic cells from yeast to mammals (1, 2). The main form of autophagy, macroautophagy, involves the non-selective sequestration of large portions of the cytoplasm into double-membrane structures termed autophagosomes, and their delivery to the vacuole/lysosome for degradation. Another process, microautophagy, involves the direct sequestration of parts of the cytoplasm by vacuole/lysosomes. The two processes coexist in yeast cells but their extent may depend on different factors including metabolic state: for example, we have observed that nitrogen-starved lactate-grown yeast cells develop microautophagy, whereas nitrogen-starved glucose-grown cells preferentially develop macroautophagy (3).Both macroautophagy and microautophagy are essentially non-selective, in the way that autophagosomes and vacuole invaginations do not appear to discriminate the sequestered material. However, selective forms of autophagy have been observed (4) that target namely peroxisomes (5, 6), chromatin (7, 8), endoplasmic reticulum (9), ribosomes (10), and mitochondria (3, 1113). Although non-selective autophagy plays an essential role in survival by nitrogen starvation, by providing amino acids to the cell, selective autophagy is more likely to have a function in the maintenance of cellular structures, both under normal conditions as a “housecleaning” process, and under stress conditions by eliminating altered organelles and macromolecular structures (1416). Selective autophagy targeting mitochondria, termed mitophagy, may be particularly relevant to stress conditions. The mitochondrial respiratory chain is both the main site and target of ROS4 production (17). Consequently, the maintenance of a pool of healthy mitochondria is a crucial challenge for the cells. The progressive accumulation of altered mitochondria (18) caused by the loss of efficiency of the maintenance process (degradation/biogenesis de novo) is often considered as a major cause of cellular aging (1923). In mammalian cells, autophagic removal of mitochondria has been shown to be triggered following induction/blockade of apoptosis (23), suggesting that autophagy of mitochondria was required for cell survival following mitochondria injury (14). Consistent with this idea, a direct alteration of mitochondrial permeability properties has been shown to induce mitochondrial autophagy (13, 24, 25). Furthermore, inactivation of catalase induced the autophagic elimination of altered mitochondria (26). In the yeast Saccharomyces cerevisiae, the alteration of F0F1-ATPase biogenesis in a conditional mutant has been shown to trigger autophagy (27). Alterations of mitochondrial ion homeostasis caused by the inactivation of the K+/H+ exchanger was shown to cause both autophagy and mitophagy (28). We have reported that treatment of cells with rapamycin induced early ROS production and mitochondrial lipid oxidation that could be inhibited by the hydrophobic antioxidant resveratrol (29). Furthermore, resveratrol treatment impaired autophagic degradation of both cytosolic and mitochondrial proteins and delayed rapamycin-induced cell death, suggesting that mitochondrial oxidation events may play a crucial role in the regulation of autophagy. This existence of regulation of autophagy by ROS has received molecular support in HeLa cells (30): these authors showed that starvation stimulated ROS production, namely H2O2, which was essential for autophagy. Furthermore, they identified the cysteine protease hsAtg4 as a direct target for oxidation by H2O2. This provided a possible connection between the mitochondrial status and regulation of autophagy.Investigations of mitochondrial autophagy in nitrogen-starved lactate-grown yeast cells have established the existence of two distinct processes: the first one occurring very early, is selective for mitochondria and is dependent on the presence of the mitochondrial protein Uth1p; the second one occurring later, is not selective for mitochondria, is not dependent on Uth1p, and is a form of bulk microautophagy (3). The absence of the selective process in the Δuth1 mutant strongly delays and decreases mitochondrial protein degradation (3, 12). The putative protein phosphatase Aup1p has been also shown to be essential in inducing mitophagy (31). Additionally several Atg proteins were shown to be involved in vacuolar sequestration of mitochondrial GFP (3, 12, 32, 33). Recently, the protein Atg11p, which had been already identified as an essential protein for selective autophagy has also been reported as being essential for mitophagy (33).The question remains as to identify of the signals that trigger selective mitophagy. It is particularly intriguing that selective mitophagy is activated very early after the shift to a nitrogen-deprived medium (3). Furthermore, selective mitophagy is very active on lactate-grown cells (with fully differentiated mitochondria) but is nearly absent in glucose-grown cells (3). In the present paper, we investigated the relationships between the redox status of the cells and selective mitophagy, namely by manipulating glutathione. Our results support the view that redox imbalance is a trigger for the selective elimination of mitochondria.  相似文献   

16.
The kinetochore, which consists of DNA sequence elements and structural proteins, is essential for high-fidelity chromosome transmission during cell division. In budding yeast, Sgt1 and Hsp90 help assemble the core kinetochore complex CBF3 by activating the CBF3 components Skp1 and Ctf13. In this study, we show that Sgt1 forms homodimers by performing in vitro and in vivo immunoprecipitation and analytical ultracentrifugation analyses. Analyses of the dimerization of Sgt1 deletion proteins showed that the Skp1-binding domain (amino acids 1–211) contains the Sgt1 homodimerization domain. Also, the Sgt1 mutant proteins that were unable to dimerize also did not bind Skp1, suggesting that Sgt1 dimerization is important for Sgt1-Skp1 binding. Restoring dimerization activity of a dimerization-deficient sgt1 mutant (sgt1-L31P) by using the CENP-B (centromere protein-B) dimerization domain suppressed the temperature sensitivity, the benomyl sensitivity, and the chromosome missegregation phenotype of sgt1-L31P. These results strongly suggest that Sgt1 dimerization is required for kinetochore assembly.Spindle microtubules are coupled to the centromeric region of the chromosome by a structural protein complex called the kinetochore (1, 2). The kinetochore is thought to generate a signal that arrests cells during mitosis when it is not properly attached to microtubules, thereby preventing aberrant chromosome transmission to the daughter cells, which can lead to tumorigenesis (3, 4). The kinetochore of the budding yeast Saccharomyces cerevisiae has been characterized thoroughly, genetically and biochemically; thus, its molecular structure is the most well detailed to date. More than 70 different proteins comprise the budding yeast kinetochore, and several of those are conserved in mammals (2).The budding yeast centromere DNA is a 125-bp region that contains three conserved regions, CDEI, CDEII, and CDEIII (5, 6). CDEI is bound by Cbf1 (79). CDEIII (25 bp) is essential for centromere function (10) and is the site where CBF3 binds to centromeric DNA. CBF3 contains four proteins: Ndc10, Cep3, Ctf13 (1118), and Skp1 (17, 18), all of which are essential for viability. Mutations in any of the four CBF3 proteins abolish the ability of CDEIII to bind to CBF3 (19, 20). All of the described kinetochore proteins, except the CDEI-binding Cbf1, localize to kinetochores dependent on the CBF3 complex (2). Therefore, the CBF3 complex is the fundamental structure of the kinetochore, and the mechanism of CBF3 assembly is of major interest.We previously isolated SGT1, the skp1-4 kinetochore-defective mutant dosage suppressor (21). Sgt1 and Skp1 activate Ctf13; thus, they are required for assembly of the CBF3 complex (21). The molecular chaperone Hsp90 is also required for the formation of the Skp1-Ctf13 complex (22). Sgt1 has two highly conserved motifs that are required for protein-protein interaction, the tetratricopeptide repeat (TPR)2 (21) and the CS (CHORD protein- and Sgt1-specific) motif. We and others (2326) have found that both domains are important for the interaction with Hsp90. The Sgt1-Hsp90 interaction is required for the assembly of the core kinetochore complex; this interaction is an initial step in kinetochore assembly (24, 26, 27) that is conserved between yeast and humans (28, 29).In this study, we further characterized the molecular mechanism of this assembly process. We found that Sgt1 forms dimers in vivo, and our results strongly suggest that Sgt1 dimerization is required for kinetochore assembly in budding yeast.  相似文献   

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

18.
Intersectin-short (intersectin-s) is a multimodule scaffolding protein functioning in constitutive and regulated forms of endocytosis in non-neuronal cells and in synaptic vesicle (SV) recycling at the neuromuscular junction of Drosophila and Caenorhabditis elegans. In vertebrates, alternative splicing generates a second isoform, intersectin-long (intersectin-l), that contains additional modular domains providing a guanine nucleotide exchange factor activity for Cdc42. In mammals, intersectin-s is expressed in multiple tissues and cells, including glia, but excluded from neurons, whereas intersectin-l is a neuron-specific isoform. Thus, intersectin-I may regulate multiple forms of endocytosis in mammalian neurons, including SV endocytosis. We now report, however, that intersectin-l is localized to somatodendritic regions of cultured hippocampal neurons, with some juxtanuclear accumulation, but is excluded from synaptophysin-labeled axon terminals. Consistently, intersectin-l knockdown (KD) does not affect SV recycling. Instead intersectin-l co-localizes with clathrin heavy chain and adaptor protein 2 in the somatodendritic region of neurons, and its KD reduces the rate of transferrin endocytosis. The protein also co-localizes with F-actin at dendritic spines, and intersectin-l KD disrupts spine maturation during development. Our data indicate that intersectin-l is indeed an important regulator of constitutive endocytosis and neuronal development but that it is not a prominent player in the regulated endocytosis of SVs.Clathrin-mediated endocytosis (CME)4 is a major mechanism by which cells take up nutrients, control the surface levels of multiple proteins, including ion channels and transporters, and regulate the coupling of signaling receptors to downstream signaling cascades (1-5). In neurons, CME takes on additional specialized roles; it is an important process regulating synaptic vesicle (SV) availability through endocytosis and recycling of SV membranes (6, 7), it shapes synaptic plasticity (8-10), and it is crucial in maintaining synaptic membranes and membrane structure (11).Numerous endocytic accessory proteins participate in CME, interacting with each other and with core components of the endocytic machinery such as clathrin heavy chain (CHC) and adaptor protein-2 (AP-2) through specific modules and peptide motifs (12). One such module is the Eps15 homology domain that binds to proteins bearing NPF motifs (13, 14). Another is the Src homology 3 (SH3) domain, which binds to proline-rich domains in protein partners (15). Intersectin is a multimodule scaffolding protein that interacts with a wide range of proteins, including several involved in CME (16). Intersectin has two N-terminal Eps15 homology domains that are responsible for binding to epsin, SCAMP1, and numb (17-19), a central coil-coiled domain that interacts with Eps15 and SNAP-23 and -25 (17, 20, 21), and five SH3 domains in its C-terminal region that interact with multiple proline-rich domain proteins, including synaptojanin, dynamin, N-WASP, CdGAP, and mSOS (16, 22-25). The rich binding capability of intersectin has linked it to various functions from CME (17, 26, 27) and signaling (22, 28, 29) to mitogenesis (30, 31) and regulation of the actin cytoskeleton (23).Intersectin functions in SV recycling at the neuromuscular junction of Drosophila and C. elegans where it acts as a scaffold, regulating the synaptic levels of endocytic accessory proteins (21, 32-34). In vertebrates, the intersectin gene is subject to alternative splicing, and a longer isoform (intersectin-l) is generated that is expressed exclusively in neurons (26, 28, 35, 36). This isoform has all the binding modules of its short (intersectin-s) counterpart but also has additional domains: a DH and a PH domain that provide guanine nucleotide exchange factor (GEF) activity specific for Cdc42 (23, 37) and a C2 domain at the C terminus. Through its GEF activity and binding to actin regulatory proteins, including N-WASP, intersectin-l has been implicated in actin regulation and the development of dendritic spines (19, 23, 24). In addition, because the rest of the binding modules are shared between intersectin-s and -l, it is generally thought that the two intersectin isoforms have the same endocytic functions. In particular, given the well defined role for the invertebrate orthologs of intersectin-s in SV endocytosis, it is thought that intersectin-l performs this role in mammalian neurons, which lack intersectin-s. Defining the complement of intersectin functional activities in mammalian neurons is particularly relevant given that the protein is involved in the pathophysiology of Down syndrome (DS). Specifically, the intersectin gene is localized on chromosome 21q22.2 and is overexpressed in DS brains (38). Interestingly, alterations in endosomal pathways are a hallmark of DS neurons and neurons from the partial trisomy 16 mouse, Ts65Dn, a model for DS (39, 40). Thus, an endocytic trafficking defect may contribute to the DS disease process.Here, the functional roles of intersectin-l were studied in cultured hippocampal neurons. We find that intersectin-l is localized to the somatodendritic regions of neurons, where it co-localizes with CHC and AP-2 and regulates the uptake of transferrin. Intersectin-l also co-localizes with actin at dendritic spines and disrupting intersectin-l function alters dendritic spine development. In contrast, intersectin-l is absent from presynaptic terminals and has little or no role in SV recycling.  相似文献   

19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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