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 62 毫秒
1.
2.
Upon DNA replication stress, stalled DNA replication forks serve as a platform to recruit many signaling proteins, leading to the activation of the DNA replication checkpoint. Activation of Rad53, a key effector kinase in the budding yeast Saccharomyces cerevisiae, is essential for stabilizing DNA replication forks during replication stress. Using an activity-based assay for Rad53, we found that Mrc1, a replication fork-associated protein, cooperates with Mec1 to activate Rad53 directly. Reconstitution of Rad53 activation using purified Mec1 and Mrc1 showed that the addition of Mrc1 stimulated a more than 70-fold increase in the ability of Mec1 to activate Rad53. Instead of increasing the catalytic activity of Mec1, Mrc1 was found to facilitate the phosphorylation of Rad53 by Mec1 via promotion of a stronger enzyme-substrate interaction between them. Further, the conserved C-terminal domain of Mrc1 was found to be required for Rad53 activation. These results thus provide insights into the role of the adaptor protein Mrc1 in activating Rad53 in the DNA replication checkpoint.Faithful replication of the genome is important for the survival of all organisms. During DNA replication, replication stress can arise from a variety of situations, including intrinsic errors made by DNA polymerases, difficulties in replicating repeated DNA sequences, and failures to repair damaged DNA caused by either endogenous oxidative agents or exogenous mutagens such as UV light and DNA-damaging chemicals (13). In eukaryotes, there is an evolutionarily conserved DNA replication checkpoint that becomes activated in response to DNA replication stress. It helps to stabilize DNA replication forks, block late replication origin firing, and delay mitosis and ultimately helps recovery from stalled replication forks after DNA repair (47). Defects in the DNA replication checkpoint could result in elevated genomic instabilities, cancer development, or cell death (8, 9).Aside from replicating the genome, the DNA replication forks also provide a platform to assemble many signaling proteins that function in the DNA replication checkpoint. In the budding yeast Saccharomyces cerevisiae, Mec1, an ortholog of human ATR,2 is a phosphoinositide 3-kinase-like kinase (PIKK) involved in sensing stalled DNA replication forks. Mec1 forms a protein complex with Ddc2 (ortholog of human ATRIP). The Mec1-Ddc2 complex is recruited to stalled replication forks through replication protein A (RPA)-coated single-stranded DNA (10, 11). The Mec3-Rad17-Ddc1 complex, a proliferating cell nuclear antigen (PCNA)-like checkpoint clamp and ortholog of the human 9-1-1 complex, was shown to be loaded onto the single- and double-stranded DNA junction of the stalled replication forks by the clamp loader Rad24-RFC complex (12). Once loaded, the Mec3-Rad17-Ddc1 complex stimulates Mec1 kinase activity (13). Dbp11 and its homolog TopBP1 in vertebrates are known components of the replication machinery (14). In addition to regulating the initiation of DNA replication, they were found to play a role in the DNA replication checkpoint (1517). They interact with the 9-1-1 complex and directly stimulate Mec1/ATR activity in vitro (1820). Thus, the assembly of multiple protein complexes at stalled DNA replication forks appears to facilitate activation of the DNA replication checkpoint (13, 18).Mrc1 (for mediator of replication checkpoint) was originally identified to be important for cells to respond to hydroxyurea in S. cerevisiae and Schizosaccharomyces pombe (21, 22). Mrc1 is a component of the DNA replisome and travels with the replication forks along chromosome during DNA synthesis (2325). Deletion of MRC1 causes defects in DNA replication, indicating its role in the normal progression of DNA replication (23). Interestingly, when DNA replication is blocked by hydroxyurea, Mrc1 undergoes Mec1- and Rad3 (S. pombe ortholog of Mec1)-dependent phosphorylation (21, 22). In S. cerevisiae, mutations of Mrc1 at the (S/T)Q sites, which are consensus phosphorylation sites of the Mec1/ATR family kinases, abolishes hydroxyurea-induced Mrc1 phosphorylation in vivo, suggesting a direct phosphorylation of Mrc1 by Mec1 (21, 22).Rad53 and Cds1, homologs of human Chk2, are the major effector kinases in the DNA replication checkpoints in S. cerevisiae and S. pombe, respectively. Activation of Rad53 is a hallmark of DNA replication checkpoint activation and is important for the maintenance of DNA replication forks in response to DNA replication stress (5, 6). Thus, it is important to understand how Rad53 activity is controlled. Interestingly, mutation of all the (S/T)Q sites of Mrc1 not only abolishes the phosphorylation of Mrc1 by Mec1 but also compromises hydroxyurea-induced Rad53 activation in S. cerevisiae (21). Similarly, mutation of the TQ sites of Mrc1 in S. pombe was shown to abolish the binding between Cds1 and Mrc1 as well as Cds1 activation (22). Further, mutation of specific TQ sites of Mrc1 in S. pombe abolishes its binding to Cds1 in vitro and the activation of Cds1 in vivo (26). Thus, Mec1/Rad3-dependent phosphorylation of Mrc1 is responsible for Mrc1 binding to Rad53/Cds1, which is essential for Rad53/Cds1 activation.An intriguing property of the Chk2 family kinases is their ability to undergo autophosphorylation and activation in the absence of other proteins in vitro (27, 28). First, autophosphorylation of a conserved threonine residue in the activation loop of Chk2 family kinase was found to be an essential part of their activation processes (26, 2931). Second, a direct and trans-phosphorylation of the N-terminal TQ sites of the Chk2 family kinases by the Mec1/ATR family kinases is also important for their activation in vivo. Analogous to the requirement of N-terminal TQ site phosphorylation of Chk2 by ATR in human (32), the activation of Rad53/Cds1 in vivo requires phosphorylation of TQ sites in their N termini by Mec1/Rad3 (33, 34).Considering that Mec1, Mrc1, and many other proteins are recruited at stalled DNA replication forks and have been shown to be involved in DNA replication checkpoint activation, a key question remains unresolved: what is the minimal system that is capable of activating Rad53 directly? Given the direct physical interaction between Mrc1 and Rad53 and the requirement of Mrc1 and Mec1 in vivo, it is likely that they both play a role in Rad53 activation. Furthermore, what is the molecular mechanism of Rad53 activation by its upstream activators? To address these questions, a faithful reconstitution of the activation of Rad53 using purified proteins is necessary. In this study, we developed an activity-based assay consisting of the Dun1 kinase, a downstream substrate of Rad53, and Sml1, as a substrate of Dun1, to quantitatively measure the activity of Rad53. Using this coupled kinase assay from Rad53 to Dun1 and then to Sml1, we screened for Mrc1 and its associated factors to see whether they could directly activate Rad53 in vitro. Our results showed that Mec1 and Mrc1 collaborate to constitute a minimal system in direct activation of Rad53.  相似文献   

3.
Collapsin response mediator protein 2 (CRMP2) is an intracellular protein that mediates signaling of Semaphorin3A (Sema3A), a repulsive axon guidance molecule. Fyn, a Src-type tyrosine kinase, is involved in the Sema3A signaling. However, the relationship between CRMP2 and Fyn in this signaling pathway is still unknown. In our research, we demonstrated that Fyn phosphorylated CRMP2 at Tyr32 residues in HEK293T cells. Immunohistochemical analysis using a phospho-specific antibody at Tyr32 of CRMP showed that Tyr32-phosphorylated CRMP was abundant in the nervous system, including dorsal root ganglion neurons, the molecular and Purkinje cell layer of adult cerebellum, and hippocampal fimbria. Overexpression of a nonphosphorylated mutant (Tyr32 to Phe32) of CRMP2 in dorsal root ganglion neurons interfered with Sema3A-induced growth cone collapse response. These results suggest that Fyn-dependent phosphorylation of CRMP2 at Tyr32 is involved in Sema3A signaling.Collapsin response mediator proteins (CRMPs)4 have been identified as intracellular proteins that mediate Semaphorin3A (Sema3A) signaling in the nervous system (1). CRMP2 is one of the five members of the CRMP family. CRMPs also mediate signal transduction of NT3, Ephrin, and Reelin (24). CRMPs interact with several intracellular molecules, including tubulin, Numb, kinesin1, and Sra1 (58). CRMPs are involved in axon guidance, axonal elongation, cell migration, synapse maturation, and the generation of neuronal polarity (1, 2, 4, 5).CRMP family proteins are known to be the major phosphoproteins in the developing brain (1, 9). CRMP2 is phosphorylated by several Ser/Thr kinases, such as Rho kinase, cyclin-dependent kinase 5 (Cdk5), and glycogen synthase kinase 3β (GSK3β) (2, 1013). The phosphorylation sites of CRMP2 by these kinases are clustered in the C terminus and have already been identified. Rho kinase phosphorylates CRMP2 at Thr555 (10). Cdk5 phosphorylates CRMP2 at Ser522, and this phosphorylation is essential for sequential phosphorylations by GSK3β at Ser518, Thr514, and Thr509 (2, 1113). These phosphorylations disrupt the interaction of CRMP2 with tubulin or Numb (2, 3, 13). The sequential phosphorylation of CRMP2 by Cdk5 and GSK3β is an essential step in Sema3A signaling (11, 13). Furthermore, the neurofibrillary tangles in the brains of people with Alzheimer disease contain hyperphosphorylated CRMP2 at Thr509, Ser518, and Ser522 (14, 15).CRMPs are also substrates of several tyrosine kinases. The phosphorylation of CRMP2 by Fes/Fps and Fer has been shown to be involved in Sema3A signaling (16, 17). Phosphorylation of CRMP2 at Tyr479 by a Src family tyrosine kinase Yes regulates CXCL12-induced T lymphocyte migration (18). We reported previously that Fyn is involved in Sema3A signaling (19). Fyn associates with PlexinA2, one of the components of the Sema3A receptor complex. Fyn also activates Cdk5 through the phosphorylation at Tyr15 of Cdk5 (19). In dorsal root ganglion (DRG) neurons from fyn-deficient mice, Sema3A-induced growth cone collapse response is attenuated compared with control mice (19). Furthermore, we recently found that Fyn phosphorylates CRMP1 and that this phosphorylation is involved in Reelin signaling (4). Although it has been shown that CRMP2 is involved in Sema3A signaling (1, 11, 13), the relationship between Fyn and CRMP2 in Sema3A signaling and the tyrosine phosphorylation site(s) of CRMPs remain unknown.Here, we show that Fyn phosphorylates CRMP2 at Tyr32. Using a phospho-specific antibody against Tyr32, we determined that the residue is phosphorylated in vivo. A nonphosphorylated mutant CRMP2Y32F inhibits Sema3A-induced growth cone collapse. These results indicate that tyrosine phosphorylation by Fyn at Tyr32 is involved in Sema3A signaling.  相似文献   

4.
5.
6.
Phosphorylation is a major post-translational modification widely used in the regulation of many cellular processes. Cyclin-dependent kinase 5 (Cdk5) is a proline-directed serine/threonine kinase activated by activation subunit p35. Cdk5-p35 regulates various neuronal activities such as neuronal migration, spine formation, synaptic activity, and cell death. The kinase activity of Cdk5 is regulated by proteolysis of p35: proteasomal degradation causes down-regulation of Cdk5, whereas cleavage of p35 by calpain causes overactivation of Cdk5. Phosphorylation of p35 determines the proteolytic pathway. We have previously identified Ser8 and Thr138 as major phosphorylation sites using metabolic labeling of cultured cells followed by two-dimensional phosphopeptide mapping and phosphospecific antibodies. However, these approaches cannot determine the extent of p35 phosphorylation in vivo. Here we report the use of Phos-tag SDS-PAGE to reveal the phosphorylation states of p35 in neuronal culture and brain. Using Phos-tag acrylamide, the electrophoretic mobility of phosphorylated p35 was delayed because it is trapped at Phos-tag sites. We found a novel phosphorylation site at Ser91, which was phosphorylated by Ca2+-calmodulin-dependent protein kinase II in vitro. We constructed phosphorylation-dependent banding profiles of p35 and Ala substitution mutants at phosphorylation sites co-expressed with Cdk5 in COS-7 cells. Using the standard banding profiles, we assigned respective bands of endogenous p35 with combinations of phosphorylation states and quantified Ser8, Ser91, and Thr138 phosphorylation. The highest level of p35 phosphorylation was observed in embryonic brain; Ser8 was phosphorylated in all p35 molecules, whereas Ser91 was phosphorylated in 60% and Thr138 was phosphorylated in ∼12% of p35 molecules. These are the first quantitative and site-specific measurements of phosphorylation of p35, demonstrating the usefulness of Phos-tag SDS-PAGE for analysis of phosphorylation states of in vivo proteins.Phosphorylation is a major post-translational modification of proteins, modulating a variety of cellular functions (1, 2). Because most phosphorylation occurs in a highly site-specific manner, identification of phosphorylation sites has been a subject of intense investigation. Several analytical methods have been utilized to identify phosphorylation sites, including mass spectrometry, amino acid sequencing, and radioisotope phosphate labeling of proteins with mutation(s) at putative phosphorylation site(s) (3, 4). Phosphorylation site-specific antibodies are frequently used to detect phosphorylation at target sites (5, 6). Many phosphospecific antibodies are now commercially available. These phosphospecific antibodies are convenient and useful tools for examining site-specific phosphorylation both in vivo and in vitro. However, they are not appropriate for estimating quantitative ratios of phosphorylation states. Electrophoretic mobility shift on SDS-PAGE is also often used to observe phosphorylation (710), but this method is not always applied to site-specific phosphorylation.Phos-tag is a newly developed dinuclear metal complex that can be used to provide phosphate-binding sites when conjugated to analytical materials such as acrylamide and biotin (11). In SDS-PAGE using Phos-tag acrylamide, phosphorylated proteins are trapped by the Phos-tag sites, delaying their migration and thus separating them from unphosphorylated proteins. Subsequent immunoblot analysis with phosphorylation-independent antibodies reveals both the phosphorylated and unphosphorylated bands. Because the migration of the phosphorylated proteins is greatly delayed compared with migration in Laemmli SDS-PAGE, it is easy to identify the phosphorylated proteins from observed positions on blots. In the past 3 years, this method has been used to detect phosphorylation states for many proteins such as ERK1/2, cdc37, myosin light chain, eIF2α, protein kinase D, β-casein, SIRT7, and dysbindin-1 (1221).Cyclin-dependent kinase 5 (Cdk5)1 is a proline-directed serine/threonine kinase that is expressed predominantly in postmitotic neurons and regulates various neuronal events such as neuronal migration, spine formation, synaptic activity, and cell death (2224). Cdk5 is activated by binding to activation subunit p35 and inactivated by proteasomal degradation of p35 (25). In addition, Cdk5 activity is deregulated by cleavage of p35 to p25 with calpain, resulting in abnormal activation and ultimately causing neuronal cell death (2629). Proteolysis of p35, either by proteasomal degradation or cleavage by calpain, is regulated by phosphorylation of p35 by Cdk5 (3033). Therefore, phosphorylation of p35 is essential for proper regulation of Cdk5 activity and function. We previously identified Ser8 and Thr138 as major p35 phosphorylation sites (33). We also showed that phosphorylation of p35 decreased during brain development and proposed its relationship to age-dependent vulnerability of neurons to stress stimuli (32). Thus, to understand the in vivo regulation of Cdk5 activity, it is critical to analyze the phosphorylation states of p35 in brain. However, there is no convenient method to analyze the precise in vivo phosphorylation status of the endogenous proteins.In this study, we applied the Phos-tag SDS-PAGE method to analyze the phosphorylation states of p35 in vivo and in cultured neurons. We constructed standard band profiles of phosphorylated p35 by Phos-tag SDS-PAGE using Ala mutants at Ser8 and/or Thr138. From these experiments, we observed an unidentified in vivo phosphorylation site at Ser91. We quantified the phosphorylation at each site in cultured neurons and brain, providing the first quantitative estimate of the in vivo phosphorylation states of p35. We discuss the usefulness of Phos-tag SDS-PAGE to analyze the in vivo phosphorylation states of proteins.  相似文献   

7.
Inflammasomes have been extensively characterized in monocytes and macrophages, but not in epithelial cells, which are the preferred host cells for many pathogens. Here we show that cervical epithelial cells express a functional inflammasome. Infection of the cells by Chlamydia trachomatis leads to activation of caspase-1, through a process requiring the NOD-like receptor family member NLRP3 and the inflammasome adaptor protein ASC. Secretion of newly synthesized virulence proteins from the chlamydial vacuole through a type III secretion apparatus results in efflux of K+ through glibenclamide-sensitive K+ channels, which in turn stimulates production of reactive oxygen species. Elevated levels of reactive oxygen species are responsible for NLRP3-dependent caspase-1 activation in the infected cells. In monocytes and macrophages, caspase-1 is involved in processing and secretion of pro-inflammatory cytokines such as interleukin-1β. However, in epithelial cells, which are not known to secrete large quantities of interleukin-1β, caspase-1 has been shown previously to enhance lipid metabolism. Here we show that, in cervical epithelial cells, caspase-1 activation is required for optimal growth of the intracellular chlamydiae.Chlamydia trachomatis is the most common cause of bacterial sexually transmitted disease in the United States, and it is the leading cause of preventable blindness in the world (15). Untreated, C. trachomatis infection in women can cause pelvic inflammatory disease, which can lead to infertility and ectopic pregnancy because of scarring of the ovaries and the Fallopian tubes (6). Infection by the lymphogranuloma venereum (LGV)2 strain of C. trachomatis, which has become more common in North America and Europe (7, 8), is characterized by swelling and inflammation of the lymph nodes in the groin (9).Chlamydiae are intracellular pathogens that preferentially infect epithelial mucosa and have a biphasic infection cycle (10). A metabolically inactive form, the elementary body, infects the epithelial host cells through entry vesicles that avoid fusion with host cell lysosomes and develop into a membrane-bound inclusion (1113). Despite their intravacuolar localization, chlamydiae are still able to acquire nutrients from the host cell and interact with host-cell signaling pathways (1323). Within a few hours, the elementary bodies differentiate into larger, metabolically active reticulate bodies, which proliferate but are noninfectious. Depending on the strain of C. trachomatis, the reticulate bodies transform back into elementary bodies after 1–3 days and are released into the extracellular medium to infect other cells (11, 24, 25). Chlamydial species possess a type III secretion (T3S) system that secretes bacterial virulence factors into host cell cytosol and may control interactions between the inclusion and host-cell compartments (26).Long before the adaptive immune response is activated, infected epithelial cells produce proinflammatory cytokines and chemokines, including interleukin (IL)-6, IL-8, and granulocyte-macrophage colony-stimulating factor (27), which recruit neutrophils to the site of infection and activate other immune effector cells. However, in many cases the immune system fails to clear the infection, and the chronic release of cytokines becomes a major contributor to the scarring and damage associated with the infection (2830).The innate immune response during C. trachomatis infection is initiated by chlamydial pathogen-associated molecular patterns, including lipopolysaccharides, which bind to pattern recognition receptors such as Toll-like receptors and cytosolic NOD-like receptors (NLRs), ultimately promoting pro-inflammatory cytokine gene expression and secretion of the cytokine proteins (3137). However, secretion of the key pro-inflammatory cytokine IL-1β is tightly regulated (38). First, pro-IL-1β is produced following activation of pattern recognition receptor, and the precursor is then cleaved into the mature form by the pro-inflammatory cysteine protease, caspase-1 (also known as interleukin-1 converting enzyme or ICE). The mechanism by which caspase-1 is activated in response to infection or tissue damage was found to be modulated by a macromolecular protein complex termed the “inflammasome,” which consists of an NLR family member, an adaptor protein (apoptosis-associated speck-like protein containing a caspase activation recruitment domain or ASC), and an inactive caspase-1 precursor (pro-caspase-1) (39, 40). Previous studies demonstrated that IL-1β is produced in response to chlamydial infection in dendritic cells, macrophages, and monocytes (4144). Moreover, C. trachomatis or Chlamydia caviae infection activates caspase-1 in epithelial cells or monocytes (43, 45, 46). However, whether caspase-1 activation during chlamydial infection requires the formation of an inflammasome remains unclear.Previous studies have shown that different pathogens can cause inflammasome-mediated caspase-1 activation in macrophages and monocytes (47). However, epithelial cells lining mucosal surfaces are not only the preferred target for chlamydial infection and other intracellular pathogens but also play an important role in early host immune response to infection by secreting proinflammatory cytokines and chemokines (27). Although epithelial cells are not known to secrete large amounts of IL-1β, inflammasome-dependent caspase-1 activation in epithelial cells is known to contribute to lipid metabolism and membrane regeneration in epithelial cells damaged by the membrane-disrupting toxin, aerolysin (48). As lipids are sorted from the Golgi apparatus to the chlamydial inclusion (13, 15, 49), we therefore investigated whether C. trachomatis induces caspase-1 activation in epithelial cells via the assembly of an inflammasome. We demonstrated that C. trachomatis-induced caspase-1 activation is mediated by an inflammasome containing the NLR member, NLRP3. Several studies have demonstrated the involvement of T3S apparatus in inflammasome-mediated caspase-1 activation by different pathogens in macrophages and monocytes (5056). Therefore, we further investigated the mechanism by which C. trachomatis triggers the formation of the NLRP3 inflammasome. Our results showed that metabolically active chlamydiae, relying on their T3S apparatus, cause K+ efflux, which in turn leads to formation of reactive oxygen species (ROS) and ultimately NLRP3-dependent caspase-1 activation. Epithelial cells do not typically secrete large amounts of IL-1β; instead, caspase-1 activation in cervical epithelial cells contributes to development of the chlamydial inclusion.  相似文献   

8.
PTEN (phosphatase and tensin homolog deleted on chromosome 10) is well characterized for its role in antagonizing the phosphoinositide 3-kinase pathway. Previous studies using size-exclusion chromatography demonstrated PTEN recruitment into high molecular mass complexes and hypothesized that PTEN phosphorylation status and PDZ binding domain may be required for such complex formation. In this study, we set out to test the structural requirements for PTEN complex assembly and identify the component(s) of the PTEN complex(es). Our results demonstrated that the PTEN catalytic function and PDZ binding domain are not absolutely required for its complex formation. On the other hand, PTEN phosphorylation status has a significant impact on its complex assembly. Our results further demonstrate enrichment of the PTEN complex in nuclear lysates, suggesting a mechanism through which PTEN phosphorylation may regulate its complex assembly. These results prompted further characterization of other protein components within the PTEN complex(es). Using size-exclusion chromatography and two-dimensional difference gel electrophoresis followed by mass spectrometry analysis, we identified heterogeneous nuclear ribonucleoprotein C (hnRNP C) as a novel protein recruited to higher molecular mass fractions in the presence of PTEN. Further analysis indicates that endogenous hnRNP C and PTEN interact and co-localize within the nucleus, suggesting a potential role for PTEN, alongside hnRNP C, in RNA regulation.Phosphatase and tensin homolog deleted on chromosome 10 (PTEN)4 was cloned in 1997 (13) and has been well characterized for its tumor-suppressive role by dephosphorylating phosphatidylinositol 3,4,5-trisphosphate to phosphatidylinositol 4,5-bisphosphate and antagonizing the phosphoinositide 3-kinase pathway (47). PTEN also regulates cell migration, cell cycle progression, DNA damage response, and chromosome stability independently of its lipid phosphatase activity through its potential protein phosphatase activity and/or protein-protein interaction (811) (for recent reviews, see 1214).PTEN is composed of an N-terminal catalytic domain and a C-terminal regulatory domain. The catalytic domain contains a conserved signature motif (HCXXGXXR) found in dual-specific protein phosphatases, and mutations within this catalytic domain, including the C124S mutation, are known to abrogate PTEN catalytic activity (4). The C terminus of PTEN contains a PDZ (PDS-95/Disc-large/Zo-1) binding domain, which interacts with PDZ-containing proteins such as MAGI-1b, MAGI-2, MAGI-3, hDLG, hMAST and NHERF (1519). In addition to the PDZ binding domain, several key serine and threonine phosphorylation sites (Ser380, Thr382, Thr383, and Ser385) at the PTEN C terminus are reported to play an important role in regulating its stability, localization, and activity (2026).Recent studies suggest that PTEN may function within higher molecular mass complexes. Through size-exclusion chromatography, Vazquez et al. (27) demonstrated that PTEN can be separated into two populations: a monomeric hyperphosphorylated subpopulation and a higher molecular mass hypophosphorylated subpopulation. It was hypothesized that PTEN in its dephosphorylated form can interact with PDZ-containing proteins such as hDLG and be recruited into a higher molecular mass complex. Although the components within PTEN complex(es) have not been systematically studied and purified, MAGI-2, hDLG (27), NHERF2, PDGFR (19), NEP (28), and MVP (29) have been identified as potential components of the PTEN complex using the same size-exclusion chromatography methodology.In this paper, we aim to (i) investigate the essential elements of PTEN required for its complex formation and (ii) dissect the components of the PTEN-associated complex(es). Our results indicate that PTEN catalytic activity or its PDZ binding domain is not absolutely required for complex assembly. PTEN phosphorylation status on amino acids Ser380, Thr382, Thr383, and Ser385, on the other hand, has a significant role in complex formation. In addition, we demonstrate that the PTEN complex is enriched in nuclear lysates, which suggests a mechanism through which phosphorylation can regulate complex assembly. Using two-dimensional difference gel electrophoresis (DIGE) analysis and comparing proteins present in higher molecular mass fractions in the presence and absence of PTEN followed by mass spectrometry analysis, we have identified heterogeneous nuclear ribonucleoprotein C (hnRNP C) as a potential component within the PTEN complex. PTEN and hnRNP C are shown here to interact and co-localize in the nucleus. We hypothesize that the PTEN and hnRNP C complex may play a role in RNA regulation.  相似文献   

9.
The human copper transporter hCTR1 is a homotrimer composed of a plasma membrane protein of 190 amino acids that contains three transmembrane segments. The extracellular 65-amino acid amino terminus of hCTR1 contains both N-linked (at Asn15) and O-linked (at Thr27) sites of glycosylation. If O-glycosylation at Thr27 is prevented, hCTR1 is efficiently cleaved, removing ∼30 amino acids from the amino terminus. We have now investigated (i) the site of this cleavage, determining which peptide bonds are cleaved, (ii) the mechanism by which glycosylation prevents cleavage, and (iii) where in the cell the proteolytic cleavage takes place. Cleavage occurs in the sequence Ala-Ser-His-Ser-His (residues 29–33), which does not contain previously recognized protease cleavage sites. Using a series of hCTR1 mutants, we show that cleavage occurs preferentially between residues Ala29–Ser30–His31. We also show that the O-linked polysaccharide at Thr27 blocks proteolysis due to its proximity to the cleavage site. Moving the cleavage site away from the Thr27 polysaccharide by insertion of as few as 5 amino acids allows cleavage to occur in the presence of glycosylation. Imaging studies using immunofluorescence in fixed cells and a functional green fluorescent protein-tagged hCTR1 transporter in live cells showed that the cleaved peptide accumulates in punctate structures in the cytoplasm. These puncta overlap compartments were stained by Rab9, indicating that hCTR1 cleavage occurs in a late endosomal compartment prior to delivery of the transporter to the plasma membrane.Copper is acquired by eukaryotic cells through transporters in the plasma membrane known as CTR proteins (1). Copper is an essential enzymatic cofactor in numerous proteins, many of which perform electron transfer reactions in which the metal cycles (2, 3) between the redox states (Cu+ and Cu2+) (4). This readily occurring redox reaction can make copper ions toxic to cells through the generation of reactive oxygen species. The free copper concentration in cells is extremely low (less than 1 fmol), and there is essentially no free copper in serum. Hence, copper transporters receive copper from copper-binding substrates in the serum, translocate it across the membrane, and transfer it to intracellular chaperones for delivery to target proteins (5).Human copper transporter 1 (hCTR1)2 and orthologous proteins throughout eukaryotes have three transmembrane segments (6, 7) and form homotrimeric, membrane complexes (8, 9) that carry out the high affinity transport of monovalent copper (see Fig. 1, inset). The human hCTR1 gene was discovered by its ability to complement Saccharomyces cerevisiae yCtr mutants, demonstrating that high affinity copper transport is a conserved function among the CTR1 proteins (10). The CTR1 proteins range in size from 200 to 400 amino acids (1, 11), but share methionine- and histidine-rich motifs in the extracellular amino terminus, as well as conserved sequences in transmembrane segments (1, 12).Open in a separate windowFIGURE 1.Extracellular amino terminus of hCTR1. Location of N- and O-linked glycosylation at Asn15 and Thr27, and the end points of 3 truncation mutants in gray: H22, A29, and G34. In the absence of O-glycosylation at Thr27, hCTR1 is efficiently cleaved between A29 and G34 (black triangles). Location of the FLAG epitope tag is shown. Inset shows the complete 190-amino acid hCTR1 protein, with extracellular NH2 terminus, three membrane spanning domains, intracellular loop, and COOH-terminal tail. The 5 amino acids in which cleavage occurs are shown in black. Three hCTR1 polypeptides form a symmetrical trimer in the copper transporter (8, 9).Little is known about the details of the copper transport mechanism in CTR1 proteins. Mutational studies of hCTR1 have identified a number of residues important for copper transport (1214), such as methionine residues within the extracellular amino terminus, and two transmembrane segments that were important for 64Cu uptake in cultured cells (12). A study of hCTR1 mutants expressed in insect cells identified residues in or near the transmembrane domains that affect Km and or Vmax of 64Cu uptake (14). These results and recent structural studies suggest that copper transits a pore lined by transmembrane segments two and three in the homotrimeric complex (8, 9). Another mechanism based on endocytosis and degradation of hCTR1 has also been proposed (15).Vertebrate CTR1 proteins are widely expressed, and may play other roles in addition to copper transport. Mice homozygous for mCtr1 knock-out alleles die during midgestation, which was thought to reflect an early requirement for copper transport during development. However, a recent study showed that xCTR1 was part of a fibroblast growth factor signaling complex in Xenopus embryos active in Ras/extracellular signal-regulated kinase (ERK) signaling. The signaling role, which affects embryonic development in Xenopus and ES cell differentiation in mammalian cells, appears to be independent of the copper transport activity of CTR1 (16).In previous structure/function studies of hCTR1 we found that the extracellular amino terminus of ∼65 amino acids is modified by N- and O-linked glycosylation at Asn15 and Thr27, respectively (6, 17) (see Fig. 1). N-Linked glycans at Asn15 increase the predicted mass of the hCTR1 polypeptide by about 9 kDa. Removing N-linked polysaccharides by a N15Q mutation does not significantly affect the expression or function of the transporter (6, 17). O-Linked polysaccharides at Thr27 that terminate in sialic acid residues increase the mass of the polypeptide by 1–2 kDa, (17). In the absence of O-linked glycosylation, the polypeptide undergoes very efficient cleavage near Thr27, leaving a 17-kDa hCTR1 protein lacking about 30 amino acids from the extracellular amino terminus (Fig. 1). The truncated (17 kDa) hCTR1 protein was efficiently delivered to the plasma membrane, but exhibited only 50–60% of the copper transport activity of wild-type hCTR1 (17).In recent years, an impressive variety of proteases have been characterized in the secretory pathway and plasma membrane (1822). Many of these proteases perform some kind of regulatory cleavage, from maturation of pre-proteins, (including proteases), to membrane proteases involved in shedding of ectodomains. Presumably, cleavage of hCTR1 lacking O-linked glycosylation must occur after the addition of the O-linked sugars would have occurred in the golgi (23). Cleavage of the unglycosylated hCTR1 protein could thus occur while the transporter is en route to the plasma membrane (23, 24), after delivery to the surface, or, as in the case of some receptors, during recycling between the plasma membrane and interior compartments (2528).In this report, we show that inhibition of cleavage by O-linked glycosylation at Thr27 requires close proximity of the polysaccharide to the site of cleavage. Moving the cleavage site away from Thr27 polysaccharides allowed cleavage. In mutants lacking O-glycosylation, hCTR1 is cleaved within amino acids 29–33 (ASHSH), preferentially between Ala29–Ser30–His31. Live cell imaging of GFP-tagged mutant hCTR1 and staining of fixed cells overexpressing FLAG-tagged hCTR1 shows that the cleaved amino-terminal peptides accumulate in punctate structures that partially overlap Rab9, a late endosome marker, suggesting that cleavage occurs after transit through the golgi, but prior to delivery to the plasma membrane.  相似文献   

10.
Human concentrative nucleoside transporter 3 (hCNT3) utilizes electrochemical gradients of both Na+ and H+ to accumulate pyrimidine and purine nucleosides within cells. We have employed radioisotope flux and electrophysiological techniques in combination with site-directed mutagenesis and heterologous expression in Xenopus oocytes to identify two conserved pore-lining glutamate residues (Glu-343 and Glu-519) with essential roles in hCNT3 Na+/nucleoside and H+/nucleoside cotransport. Mutation of Glu-343 and Glu-519 to aspartate, glutamine, and cysteine severely compromised hCNT3 transport function, and changes included altered nucleoside and cation activation kinetics (all mutants), loss or impairment of H+ dependence (all mutants), shift in Na+:nucleoside stoichiometry from 2:1 to 1:1 (E519C), complete loss of catalytic activity (E519Q) and, similar to the corresponding mutant in Na+-specific hCNT1, uncoupled Na+ currents (E343Q). Consistent with close-proximity integration of cation/solute-binding sites within a common cation/permeant translocation pore, mutation of Glu-343 and Glu-519 also altered hCNT3 nucleoside transport selectivity. Both residues were accessible to the external medium and inhibited by p-chloromercuribenzene sulfonate when converted to cysteine.Physiologic nucleosides and the majority of synthetic nucleoside analogs with antineoplastic and/or antiviral activity are hydrophilic molecules that require specialized plasma membrane nucleoside transporter (NT)3 proteins for transport into or out of cells (14). NT-mediated transport is required for nucleoside metabolism by salvage pathways and is a critical determinant of the pharmacologic actions of nucleoside drugs (36). By regulating adenosine availability to purinoreceptors, NTs also modulate a diverse array of physiological processes, including neurotransmission, immune responses, platelet aggregation, renal function, and coronary vasodilation (4, 6, 7). Two structurally unrelated NT families of integral membrane proteins exist in human and other mammalian cells and tissues as follows: the SLC28 concentrative nucleoside transporter (CNT) family and the SLC29 equilibrative nucleoside transporter (ENT) family (3, 4, 6, 8, 9). ENTs are normally present in most, possibly all, cell types (4, 6, 8). CNTs, in contrast, are found predominantly in intestinal and renal epithelia and other specialized cell types, where they have important roles in absorption, secretion, distribution, and elimination of nucleosides and nucleoside drugs (13, 5, 6, 9).The CNT protein family in humans is represented by three members, hCNT1, hCNT2, and hCNT3. Belonging to a CNT subfamily phylogenetically distinct from hCNT1/2, hCNT3 utilizes electrochemical gradients of both Na+ and H+ to accumulate a broad range of pyrimidine and purine nucleosides and nucleoside drugs within cells (10, 11). hCNT1 and hCNT2, in contrast, are Na+-specific and transport pyrimidine and purine nucleosides, respectively (1113). Together, hCNT1–3 account for the three major concentrative nucleoside transport processes of human and other mammalian cells. Nonmammalian members of the CNT protein family that have been characterized functionally include hfCNT, a second member of the CNT3 subfamily from the ancient marine prevertebrate the Pacific hagfish Eptatretus stouti (14), CeCNT3 from Caenorhabditis elegans (15), CaCNT from Candida albicans (16), and the bacterial nucleoside transporter NupC from Escherichia coli (17). hfCNT is Na+- but not H+-coupled, whereas CeCNT3, CaCNT, and NupC are exclusively H+-coupled. Na+:nucleoside coupling stoichiometries are 1:1 for hCNT1 and hCNT2 and 2:1 for hCNT3 and hfCNT3 (11, 14). H+:nucleoside coupling ratios for hCNT3 and CaCNT are 1:1 (11, 16).Although much progress has been made in molecular studies of ENT proteins (4, 6, 8), studies of structurally and functionally important regions and residues within the CNT protein family are still at an early stage. Topological investigations suggest that hCNT1–3 and other eukaryote CNT family members have a 13 (or possibly 15)-transmembrane helix (TM) architecture, and multiple alignments reveal strong sequence similarities within the C-terminal half of the proteins (18). Prokaryotic CNTs lack the first three TMs of their eukaryotic counterparts, and functional expression of N-terminally truncated human and rat CNT1 in Xenopus oocytes has established that these three TMs are not required for Na+-dependent uridine transport activity (18). Consistent with this finding, chimeric studies involving hCNT1 and hfCNT (14) and hCNT1 and hCNT3 (19) have demonstrated that residues involved in Na+- and H+-coupling reside in the C-terminal half of the protein. Present in this region of the transporter, but of unknown function, is a highly conserved (G/A)XKX3NEFVA(Y/M/F) motif common to all eukaryote and prokaryote CNTs.By virtue of their negative charge and consequent ability to interact directly with coupling cations and/or participate in cation-induced and other protein conformational transitions, glutamate and aspartate residues play key functional and structural roles in a broad spectrum of mammalian and bacterial cation-coupled transporters (2030). Little, however, is known about their role in CNTs. This study builds upon a recent mutagenesis study of conserved glutamate and aspartate residues in hCNT1 (31) to undertake a parallel in depth investigation of corresponding residues in hCNT3. By employing the multifunctional capability of hCNT3 as a template for these studies, this study provides novel mechanistic insights into the molecular mechanism(s) of CNT-mediated cation/nucleoside cotransport, including the role of the (G/A)XKX3NEFVA(Y/M/F) motif.  相似文献   

11.
The binding of the adaptor protein APPL1 to adiponectin receptors is necessary for adiponectin-induced AMP-activated protein kinase (AMPK) activation in muscle, yet the underlying molecular mechanism remains unknown. Here we show that in muscle cells adiponectin and metformin induce AMPK activation by promoting APPL1-dependent LKB1 cytosolic translocation. APPL1 mediates adiponectin signaling by directly interacting with adiponectin receptors and enhances LKB1 cytosolic localization by anchoring this kinase in the cytosol. Adiponectin also activates another AMPK upstream kinase Ca2+/calmodulin-dependent protein kinase kinase by activating phospholipase C and subsequently inducing Ca2+ release from the endoplasmic reticulum, which plays a minor role in AMPK activation. Our results show that in muscle cells adiponectin is able to activate AMPK via two distinct mechanisms as follows: a major pathway (the APPL1/LKB1-dependent pathway) that promotes the cytosolic localization of LKB1 and a minor pathway (the phospholipase C/Ca2+/Ca2+/calmodulin-dependent protein kinase kinase-dependent pathway) that stimulates Ca2+ release from intracellular stores.Adiponectin, an adipokine abundantly expressed in adipose tissue, exhibits anti-diabetic, anti-inflammatory, and anti-atherogenic properties and hence is a potential therapeutic target for various metabolic diseases (13). The beneficial effects of adiponectin are mediated through the direct interaction of adiponectin with its cell surface receptors, AdipoR1 and AdipoR2 (4, 5). Adiponectin increases fatty acid oxidation and glucose uptake in muscle cells by activating AMP-activated protein kinase (AMPK)3 (4, 6), which depends on the interaction of AdipoR1 with the adaptor protein APPL1 (Adaptor protein containing Pleckstrin homology domain, Phosphotyrosine binding domain, and Leucine zipper motif) (5). However, the underlying mechanisms by which APPL1 mediates adiponectin signaling to AMPK activation and other downstream targets remain unclear.AMPK is a serine/threonine protein kinase that acts as a master sensor of cellular energy balance in mammalian cells by regulating glucose and lipid metabolism (7, 8). AMPK is composed of a catalytic α subunit and two noncatalytic regulatory subunits, β and γ. The NH2-terminal catalytic domain of the AMPKα subunit is highly conserved and contains the activating phosphorylation site (Thr172) (9). Two AMPK variants, α1 and α2, exist in mammalian cells that show different localization patterns. AMPKα1 subunit is localized in non-nuclear fractions, whereas the AMPKα2 subunit is found in both nucleus and non-nuclear fractions (10). Biochemical regulation of AMPK activation occurs through various mechanisms. An increase in AMP level stimulates the binding of AMP to the γ subunit, which induces a conformational change in the AMPK heterotrimer and results in AMPK activation (11). Studies have shown that the increase in AMPK activity is not solely via AMP-dependent conformational change, rather via phosphorylation by upstream kinases, LKB1 and CaMKK. Dephosphorylation by protein phosphatases is also important in regulating the activity of AMPK (12).LKB1 has been considered as a constitutively active serine/threonine protein kinase that is ubiquitously expressed in all tissues (13, 14). Under conditions of high cellular energy stress, LKB1 acts as the primary AMPK kinase through an AMP-dependent mechanism (1517). Under normal physiological conditions, LKB1 is predominantly localized in the nucleus. LKB1 is translocated to the cytosol, either by forming a heterotrimeric complex with Ste20-related adaptor protein (STRADα/β) and mouse protein 25 (MO25α/β) or by associating with an LKB1-interacting protein (LIP1), to exert its biological function (1822). Although LKB1 has been shown to mediate contraction- and adiponectin-induced activation of AMPK in muscle cells, the underlying molecular mechanisms remain elusive (15, 23).CaMKK is another upstream kinase of AMPK, which shows considerable sequence and structural homology with LKB1 (2426). The two isoforms of CaMKK, CaMKKα and CaMKKβ, encoded by two distinct genes, share ∼70% homology at the amino acid sequence level and exhibit a wide expression in rodent tissues, including skeletal muscle (2734). Unlike LKB1, AMPK phosphorylation mediated by CaMKKs is independent of AMP and is dependent only on Ca2+/calmodulin (35). Hence, it is possible that an LKB1-independent activation of AMPK by CaMKK exists in muscle cells. However, whether and how adiponectin stimulates this pathway in muscle cells are not known.In this study, we demonstrate that in muscle cells adiponectin induces an APPL1-dependent LKB1 translocation from the nucleus to the cytosol, leading to increased AMPK activation. Adiponectin also activates CaMKK by stimulating intracellular Ca2+ release via the PLC-dependent mechanism, which plays a minor role in activation of AMPK. Taken together, our results demonstrate that enhanced cytosolic localization of LKB1 and Ca2+-induced activation of CaMKK are the mechanisms underlying adiponectin-stimulated AMPK activation in muscle cells.  相似文献   

12.
13.
The cell cycle checkpoint kinases play central roles in the genome maintenance of eukaryotes. Activation of the yeast checkpoint kinase Rad53 involves Rad9 or Mrc1 adaptor-mediated phospho-priming by Mec1 kinase, followed by auto-activating phosphorylation within its activation loop. However, the mechanisms by which these adaptors regulate priming phosphorylation of specific sites and how this then leads to Rad53 activation remain poorly understood. Here we used quantitative mass spectrometry to delineate the stepwise phosphorylation events in the activation of endogenous Rad53 in response to S phase alkylation DNA damage, and we show that the two Rad9 and Mrc1 adaptors, the four N-terminal Mec1-target TQ sites of Rad53 (Rad53-SCD1), and Rad53-FHA2 coordinate intimately for optimal priming phosphorylation to support substantial Rad53 auto-activation. Rad9 or Mrc1 alone can mediate surprisingly similar Mec1 target site phosphorylation patterns of Rad53, including previously undetected tri- and tetraphosphorylation of Rad53-SCD1. Reducing the number of TQ motifs turns the SCD1 into a proportionally poorer Mec1 target, which then requires the presence of both Mrc1 and Rad9 for sufficient priming and auto-activation. The phosphothreonine-interacting Rad53-FHA domains, particularly FHA2, regulate phospho-priming by interacting with the checkpoint mediators but do not seem to play a major role in the phospho-SCD1-dependent auto-activation step. Finally, mutation of all four SCD1 TQ motifs greatly reduces Rad53 activation but does not eliminate it, and residual Rad53 activity in this mutant is dependent on Rad9 but not Mrc1. Altogether, our results provide a paradigm for how phosphorylation site clusters and checkpoint mediators can be involved in the regulation of signaling relay in protein kinase cascades in vivo and elucidate an SCD1-independent Rad53 auto-activation mechanism through the Rad9 pathway. The work also demonstrates the power of mass spectrometry for in-depth analyses of molecular mechanisms in cellular signaling in vivo.Eukaryotic cells are most vulnerable to exogenous DNA-damaging agents during the S phase of the cell cycle, when unprogrammed DNA lesions interfere with the tightly choreographed DNA replication process. DNA damage during this phase leads to the activation of two overlapping checkpoint pathways in Saccharomyces cerevisiae, the DNA replication checkpoint and the intra-S-phase DNA damage checkpoint (1, 2). Phospho-priming for auto-activation of the central checkpoint kinase Rad53 by the upstream kinase Mec1/Tel1 depends on Mrc1 as an adaptor in the DNA replication checkpoint pathway and Rad9 as an adaptor in the DNA damage checkpoint pathway (310). Rad53, a well-accepted model system for studying the function and regulation of Chk2-like kinases, contains two forkhead-associated (FHA)1 domains (FHA1 and -2) and two SQ/TQ cluster domains (SCD1 and -2) enriched in Mec1/Tel1-target phosphorylation sites (1113).Mrc1 normally is a replisome component that functionally couples DNA Pol ε with Cdc45 and MCM helicase during replication fork progression (14, 15). As the replication forks are stalled by replication stress, the recruited checkpoint sensor kinase Mec1 phosphorylates the SCD of Mrc1, which abolishes its N-terminal interaction with Pol ε and enables Mrc1 to recruit Rad53 and promote Rad53 phosphorylation by Mec1 as an initial step in the activation of Rad53 in the Mrc1 branch (6, 14, 16). Alanine substitution of all Mec1 target sites of Mrc1 (designated the mrc1-AQ allele) has been shown to selectively disable its checkpoint function for Rad53 activation without affecting its DNA replication functions (4). In response to DNA damage, Rad9 is able to associate with damaged chromatin via its BRCT and Tudor domains, which tether it to Ser129-phosphorylated histone H2A (γH2A) and Lys79-methylated histone H3, respectively (17, 18). Alternatively, the recruitment of Rad9 onto damaged DNA could also be facilitated by its phosphorylation by CDK1, which enables the specific interaction of Rad9 with Dpb11, allowing the formation of the ternary complex of Dpb11, Mec1, and Rad9 (19, 20). Similar to Mrc1, Mec1 activates the adaptor function of Rad9 by phosphorylation of its SCD, which then binds to the Rad53-FHA domains to promote Rad53 phosphorylation by Mec1 (3, 5, 10).Beyond serving as scaffolds to recruit Rad53, Mrc1 and Rad9 have been shown to promote Rad53 phosphorylation by Mec1 in a dose-dependent manner in vitro (3, 16), underlining their adaptor role to enhance the enzyme–substrate (Mec1–Rad53) interaction. However, how they can specifically regulate the priming phosphorylation at specific sites and how this then leads to Rad53 activation remains poorly understood. Finally, hyperphosphorylated Rad9 has also been shown to catalyze the auto-phosphorylation of recombinant Rad53 (21), but it remains to be examined whether and how this occurs in vivo.The activation of SCD-FHA containing kinases such as human Chk2 and fission yeast Cds1 has been suggested to involve a two-step phosphorylation process: first, SCD phosphorylation by an ATM/ATR-like kinase leads to intermolecular binding to the FHA domain of another Chk2/Cds1 monomer, which then results in dimerization/oligomerization-dependent auto-phosphorylation within the kinase activation loop (2226). In addition to the characteristic N-terminal SCD-FHA module of Chk2-like kinases, Rad53 contains another SCD2-FHA2 module C-terminal to its kinase domain. Similar to its orthologues, Rad53 activation has been proposed to depend on SCD1 phosphorylation (but not SCD2 phosphorylation) and partially redundant functions of the two FHA domains (9, 2729). However, although Rad53-FHA1 can interact with SCD1 in a phospho-threonine (pT)-dependent manner in vitro (9, 28), it appears to be required for Rad53 activation only in G2/M-arrested cells (27, 29). In contrast, the FHA2 domain, which seems to be more important overall for Rad53 activation, does not appreciably bind phospho-SCD1 peptides in vitro (27, 28). Thus, the mechanisms by which Mrc1, Rad9, SCD1 phosphorylation, and FHA domains interact during checkpoint-dependent Rad53 priming and auto-activation remain to be elucidated.Quantitative mass spectrometric analysis has revolutionized the functional analysis of cellular signaling pathways, including site-specific phosphorylation events of key signaling molecules (3033), but an important caveat is that MS studies often involve protein tags or nonphysiological expression levels that can interfere with normal protein functions. For example, the integration of a triple HA tag into the endogenous RAD53 gene locus has been shown to reduce Rad53 protein levels, resulting in significantly altered checkpoint activity (34). In this study we used quantitative MS analyses to dissect the stepwise phosphorylation events of endogenous, untagged Rad53 in response to MMS-induced alkylation DNA damage and replication stress during the S phase. Together with functional analyses, our results delineate how the two Mec1 adaptors Rad9 and Mrc1 can coordinate with the four SCD1 priming sites (T5, T8, T12, and T15) to regulate the phospho-priming of Rad53 by Mec1. In addition, an SCD1-priming independent Rad53 auto-activation mechanism and the specific roles of the FHA domains during Rad53 hyperphosphorylation are also elucidated in this work.  相似文献   

14.
15.
16.
17.
Ca2+/calmodulin-dependent protein kinase II (αCaMKII) is thought to exert its role in memory formation by autonomous Ca2+-independent persistent activity conferred by Thr286 autophosphorylation, allowing the enzyme to remain active even when intracellular [Ca2+] has returned to resting levels. Ca2+ sequestration-induced inhibition, caused by a burst of Thr305/306 autophosphorylation via calmodulin (CaM) dissociation from the Thr305/306 sites, is in conflict with this view. The processes of CaM binding, autophosphorylation, and inactivation are dissected to resolve this conflict. Upon Ca2+ withdrawal, CaM sequential domain dissociation is observed, starting with the rapid release of the first (presumed N-terminal) CaM lobe, thought to be bound at the Thr305/306 sites. The time courses of Thr305/306 autophosphorylation and inactivation, however, correlate with the slow dissociation of the second (presumed C-terminal) CaM lobe. Exposure of the Thr305/306 sites is thus not sufficient for their autophosphorylation. Moreover, Thr305/306 autophosphorylation and autoinactivation are shown to occur in the continuous presence of Ca2+ and bound Ca2+/CaM by time courses similar to those seen following Ca2+ sequestration. Our investigation of the activity and mechanisms of phospho-Thr286-αCaMKII thus shows time-dependent autoinactivation, irrespective of the continued presence of Ca2+ and CaM, allowing a very short, if any, time window for Ca2+/CaM-free phospho-Thr286-αCaMKII activity. Physiologically, the time-dependent autoinactivation mechanisms of phospho-Thr286-αCaMKII (t½ of ∼50 s at 37 °C) suggest a transient kinase activity of ∼1 min duration in the induction of long term potentiation and thus memory formation.Ca2+/calmodulin-dependent protein kinase II (αCaMKII)2 is essential in hippocampal learning and N-methyl-d-aspartate receptor-dependent synaptic plasticity, causing long term potentiation (1, 2). The exact mechanisms of αCaMKII in memory functions have not yet been identified.αCaMKII is a broad specificity Ser/Thr protein kinase, which catalyzes the phosphorylation of over 100 protein and peptide substrates in vitro (3). Uniquely, the CaMKII family possesses two distinct kinase mechanisms. The first mechanism is a “canonical” intrasubunit phosphorylation, commonly found in monomeric kinases, in which the phosphorylatable residue of the substrate bound to the helical subdomain of the catalytic domain at the active site is lined up with the terminal phosphate of ATP (4). Although there is a large number of potential “canonical” substrates for αCaMKII at the synapse (5), so far AMPA receptors have been shown to be possible physiological substrates of αCaMKII (6). For the purpose of this study, syntide 2, a commonly used peptide substrate derived from phosphorylation site 2 of glycogen synthase (7), was chosen.The second mechanism, intersubunit autophosphorylation, takes advantage of the oligomeric organization of CaMKII (8). The most important autophosphorylation site in the α isoform is Thr286, which resides in the vicinity of the autoinhibitory domain (9). Peptide substrates with homologous sequences to this region have been reported to be phosphorylated by αCaMKII. This, however, occurs with a low Vmax, and these substrates show properties of a non-competitive inhibitor with respect to phosphorylation of “canonical” substrates (10) and of Thr286 autophosphorylation itself (11). Examples of such substrates include autocamtide, a peptide substrate derived from the autoinhibitory region (12) and the NR2B subunit of the N-methyl-d-aspartate receptor, which has been identified as a potential physiological target of phospho-Thr286-αCaMKII at the postsynaptic membrane (13). The possible physiological significance of NR2B phosphorylation is not yet known. There is evidence to suggest that Thr286 autophosphorylation is required to achieve full activity of the enzyme, since the unphosphorylatable T286A mutant enzyme has much diminished activity compared with wild type enzyme (14, 15).Thr286 autophosphorylation causes CaM “trapping,” a >104-fold increase in the affinity of αCaMKII for Ca2+/CaM (1618). At the same time, Thr286 autophosphorylation is also attributed to confer Ca2+- and CaM-independent persistent “autonomous” kinase activity to αCaMKII. However, due to the extremely high affinity of phospho-Thr286-αCaMKII for Ca2+/CaM, [Ca2+] of <10 nm is required to achieve full dissociation of Ca2+/CaM, since CaM trapping occurs by virtue of Ca2+ trapping (19). Partial activity measured upon partial Ca2+ withdrawal therefore may not always reflect Ca2+/CaM-free enzyme (9). Furthermore, the physiological resting [Ca2+] range is 50–100 nm; therefore, phospho-Thr286-αCaMKII is likely always to have residual Ca2+/CaM bound. This may be partially Ca2+-saturated CaM (19).Persistent autonomous activity conferred by Thr286 autophosphorylation is thought to enable αCaMKII to function as a memory molecule (20, 21). In contrast, however, following the development of chemical long term potentiation, rapid inactivation has also been reported (22). The extent of an autonomous activity is further obscured by the finding that Ca2+ sequestration induces a burst of autophosphorylation at residues Thr305/306, followed by a loss of activity (23). Moreover, when examined across a broad range of [Ca2+], the Ca2+/CaM dependence of phospho-Thr286-αCaMKII activity is apparent (19). It is thus vital to establish the mechanisms of activation and inactivation of αCaMKII at the molecular level in order to understand how it may function physiologically in learning and memory. To this end, it is necessary to dissect the mechanisms of Ca2+/CaM dissociation, Thr305/306 autophosphorylation, and inactivation of phospho-Thr286-αCaMKII and to establish the time window for autonomous Ca2+/CaM-independent activity.  相似文献   

18.
The general stress response of Bacillus subtilis can be activated by stimuli such as the addition of salt or ethanol and with blue light. In the latter response, YtvA activates σB through a cascade of Rsb proteins, organized in stressosomes. YtvA is composed of an N-terminal LOV (light, oxygen, and voltage) domain and a C-terminal STAS (sulfate transporter and anti-sigma factor) domain and shows light-modulated GTP binding in vitro. Here, we examine the mechanism of YtvA-mediated activation of σB in vivo with site-directed mutagenesis. Constitutive off and constitutive on mutations have been identified. Disruption of GTP binding in the STAS domain eliminates light activation of σB. In contrast, modification of sites relevant for phosphorylation of STAS domains does not affect the stress response significantly. The data obtained are integrated into a model for the structure of full-length YtvA, which presumably functions as a dimer.LOV2 domains (1), members of the superfamily of PAS domains (2, 3), are abundant in all domains of life and were first identified in plant phototropins (4). These photoreceptors regulate stomatal opening, phototropism, etc. and contain two N-terminal LOV domains that confer light regulation on the C-terminal Ser/Thr kinase domain (4). They also occur in bacteria, in which YtvA from Bacillus subtilis has been best characterized (for a review, see e.g. Ref. 5). Its N-terminal LOV domain binds FMN and shows the typical LOV photochemistry (6, 7): covalent adduct formation between a cysteine and the FMN chromophore. A linker helix, denoted Jα (7), connects the LOV domain to a STAS domain. The latter domain is present in many regulators of the general stress response of B. subtilis (8, 9). Stress via the addition of salt or ethanol (for a review, see Ref. 10) and blue light (11, 12) activates the general stress response via the environmental pathway, which integrates various signals via a large multiprotein complex, called the stressosome (13, 14). YtvA, which mediates light activation of σB (11, 12, 15), co-purifies with other STAS domain proteins in the stressosomes (16).When cells are stressed, STAS domains of several stressosome proteins (e.g. RsbS and RsbR) are phosphorylated by another intrinsic stressosome component, the serine/threonine kinase RsbT (9, 14, 17, 18). Next, RsbT is released from the complex to trigger RsbU, a protein phosphatase, thus (indirectly) activating σB (19). Phosphorylation of YtvA, however, has never been detected. Rather, it has been demonstrated in vitro that YtvA shows light-dependent GTP binding, presumably at its NTP-binding site in the STAS domain (20).Little is known about the mechanism of signal transmission in and by YtvA, except that in the C62A mutant, photochemistry in vitro (12) and light activation of σB in vivo (12, 15) are abolished. More detailed information is available for LOV domains of phototropins. A conserved glutamine, which is in hydrogen-bonding contact with the isoalloxazine ring of FMN, rotates its side chain by 180° upon covalent adduct formation (21). Replacement of this residue by leucine in the LOV2 domain of Phy3 from Adiantum results in a considerable reduction of the light-induced structural change (22). The corresponding mutation in phototropin 1 from Arabidopsis impairs autophosphorylation activity (23). The signal generated in the LOV2 domain is transmitted to the downstream kinase domain of phototropin 1 of Avena sativa through disruption of the interaction between its central β-sheet and the C-terminal linker region, the Jα-helix (24).Here, we study the mechanism of activation of YtvA in vivo, i.e. light-induced activation of the σB response, with site-directed mutagenesis. We focus on three regions of the protein, the flavin-binding pocket, the β-sheet of the LOV domain, and the GTP-binding site, and on potential phosphorylation sites of the STAS domain. We demonstrate that light-activated GTP binding is crucial for functional YtvA. A computational approach was used to model the structure of full-length YtvA. The model suggests that light modulates accessibility of the GTP-binding site of the STAS domain of YtvA.  相似文献   

19.
20.
Serine proteinases in insect plasma have been implicated in two types of immune responses; that is, activation of prophenoloxidase (proPO) and activation of cytokine-like proteins. We have identified more than 20 serine proteinases in hemolymph of the tobacco hornworm, Manduca sexta, but functions are known for only a few of them. We report here functions of two additional M. sexta proteinases, hemolymph proteinases 6 and 8 (HP6 and HP8). HP6 and HP8 are each composed of an amino-terminal clip domain and a carboxyl-terminal proteinase domain. HP6 is an apparent ortholog of Drosophila Persephone, whereas HP8 is most similar to Drosophila and Tenebrio spätzle-activating enzymes, all of which activate the Toll pathway. proHP6 and proHP8 are expressed constitutively in fat body and hemocytes and secreted into plasma, where they are activated by proteolytic cleavage in response to infection. To investigate activation and biological activity of HP6 and HP8, we purified recombinant proHP8, proHP6, and mutants of proHP6 in which the catalytic serine was replaced with alanine, and/or the activation site was changed to permit activation by bovine factor Xa. HP6 was found to activate proPO-activating proteinase (proPAP1) in vitro and induce proPO activation in plasma. HP6 was also determined to activate proHP8. Active HP6 or HP8 injected into larvae induced expression of antimicrobial peptides and proteins, including attacin, cecropin, gloverin, moricin, and lysozyme. Our results suggest that proHP6 becomes activated in response to microbial infection and participates in two immune pathways; activation of PAP1, which leads to proPO activation and melanin synthesis, and activation of HP8, which stimulates a Toll-like pathway.Innate immune systems of mammals and arthropods include extracellular serine proteinase cascade pathways, which rapidly amplify responses to infection and stimulate killing of pathogens. These proteinase-driven processes include the complement system of vertebrates (1, 2) and pathways in arthropods involving proteinases containing amino-terminal clip domains (3). Clip domain proteinases function in blood coagulation (4, 5), activation of prophenoloxidase (proPO) that leads to melanin synthesis (69), and stimulation of the Toll pathway to promote synthesis of antimicrobial peptides/proteins (AMPs)2 secreted into the hemolymph (10, 11).The serine proteinase systems best characterized in arthropods are the horseshoe crab hemolymph coagulation pathway and the cascade leading to activation of the Toll pathway in dorsal-ventral development in Drosophila (1214). Recent research also has led to better characterization of the proPO activation pathway in Manduca sexta (7, 15, 16) and the Toll-signaling pathway in the Drosophila immune response (17, 18) and to both the proPO and Toll pathways in the beetle Tenebrio molitor (11, 19).In the proPO activation pathway, soluble pattern recognition proteins initially recognize pathogen-associated molecular patterns such as bacterial peptidoglycan or fungal β-1,3-glucan (2022). This interaction stimulates the sequential activation of a series of serine proteinases in hemolymph, leading to the activation of proPO-activating proteinase (PAP), also known as proPO activating enzyme (7, 23). Activated PAP converts inactive proPO to PO. PO catalyzes the hydroxylation of monophenols to o-diphenols and the oxidation of o-diphenols to quinones that are involved in microbial killing, melanin synthesis, sequestration of parasites or pathogens, and wound healing (24, 25). Other proteins required for proPO activation are clip-domain serine proteinase homologs (SPHs), whose catalytic serine is replaced with glycine and, therefore, lack proteolytic activity (26, 27). Serine proteinase inhibitors, including members of the serpin superfamily, regulate the activation of proPO by inhibiting the activating proteinases (28, 29).Drosophila clip-domain serine proteinases Persephone, Grass, Spirit, and spätzle-processing enzyme (SPE) participate in the activation of Toll pathway, stimulating synthesis of antimicrobial peptides as an innate immune response (18, 3032). Although genetic evidence indicates that Persephone and Spirit are upstream of SPE in the cascade, the substrate(s) of Persephone and Spirit have not been identified, and which proteinase directly activates SPE is unknown. Neither is it clear whether these enzymes may be related to the melanization pathway, which involves clip-domain proteinases MP2 and MP1 (33).Here we report the functional characterization of M. sexta HP6 and HP8, probable orthologs of Drosophila Persephone and SPE, respectively. We developed methods to activate purified recombinant proHP6 and proHP8 and discovered that HP6 participates in proPO activation by activating proPAP1 and that both HP6 and HP8 function in a pathway that stimulates the synthesis of AMPs in M. sexta.  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