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.
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   

2.
3.
4.
Stathmin is an important regulator of microtubule polymerization and dynamics. When unphosphorylated it destabilizes microtubules in two ways, by reducing the microtubule polymer mass through sequestration of soluble tubulin into an assembly-incompetent T2S complex (two α:β tubulin dimers per molecule of stathmin), and by increasing the switching frequency (catastrophe frequency) from growth to shortening at plus and minus ends by binding directly to the microtubules. Phosphorylation of stathmin on one or more of its four serine residues (Ser16, Ser25, Ser38, and Ser63) reduces its microtubule-destabilizing activity. However, the effects of phosphorylation of the individual serine residues of stathmin on microtubule dynamic instability have not been investigated systematically. Here we analyzed the effects of stathmin singly phosphorylated at Ser16 or Ser63, and doubly phosphorylated at Ser25 and Ser38, on its ability to modulate microtubule dynamic instability at steady-state in vitro. Phosphorylation at either Ser16 or Ser63 strongly reduced or abolished the ability of stathmin to bind to and sequester soluble tubulin and its ability to act as a catastrophe factor by directly binding to the microtubules. In contrast, double phosphorylation of Ser25 and Ser38 did not affect the binding of stathmin to tubulin or microtubules or its catastrophe-promoting activity. Our results indicate that the effects of stathmin on dynamic instability are strongly but differently attenuated by phosphorylation at Ser16 and Ser63 and support the hypothesis that selective targeting by Ser16-specific or Ser63-specific kinases provides complimentary mechanisms for regulating microtubule function.Stathmin is an 18-kDa ubiquitously expressed microtubule-destabilizing phosphoprotein whose activity is modulated by phosphorylation of its four serine residues, Ser16, Ser25, Ser38, and Ser63 (17). Several classes of kinases have been identified that phosphorylate stathmin, including kinases associated with cell growth and differentiation such as members of the mitogen-activated protein kinase (MAPK)2 family, cAMP-dependent protein kinase (15, 811), and kinases associated with cell cycle regulation such as cyclin-dependent kinase 1 (3, 1214). Phosphorylation of stathmin is required for cell cycle progression through mitosis and for proper assembly/function of the mitotic spindle (3, 1316). Inhibition of stathmin phosphorylation produces strong mitotic phenotypes characterized by disassembly and disorganization of mitotic spindles and abnormal chromosome distributions (3, 1314).Stathmin is known to destabilize microtubules in two ways. One is by binding to soluble tubulin and forming a stable complex that cannot polymerize into microtubules, consisting of one molecule of stathmin and two molecules of tubulin (T2S complex) (1724). Addition of stathmin to microtubules in equilibrium with soluble tubulin results in sequestration of the tubulin and a reduction in the level of microtubule polymer (1718, 22, 2528). In addition to reducing the amount of assembled polymer, tubulin sequestration by stathmin has been shown to increase the switching frequency at microtubule plus ends from growth to shortening (called the catastrophe frequency) as the microtubules relax to a new steady state (17, 29). The second way is by binding directly to microtubules (2730). The direct binding of stathmin to microtubules increases the catastrophe frequency at both ends of the microtubules and considerably more strongly at minus ends than at plus ends (27). Consistent with its strong catastrophe-promoting activity at minus ends, stathmin increases the treadmilling rate of steady-state microtubules in vitro (27). These results have led to the suggestion that stathmin might be an important cellular regulator of minus-end microtubule dynamics (27).Phosphorylation of stathmin diminishes its ability to regulate microtubule polymerization (3, 14, 2526). Phosphorylation of Ser16 or Ser63 appears to be more critical than phosphorylation of Ser25 and Ser38 for the ability of stathmin to bind to soluble tubulin and to inhibit microtubule assembly in vitro (3, 25). Inhibition of stathmin phosphorylation induces defects in spindle assembly and organization (3, 14) suggesting that not only soluble tubulin-microtubule levels are regulated by phosphorylation of stathmin, but the dynamics of microtubules could also be regulated in a phosphorylation-dependent manner.It is not known how phosphorylation at any of the four serine residues of stathmin affects its ability to regulate microtubule dynamics and, specifically, its ability to increase the catastrophe frequency at plus and minus ends due to its direct interaction with microtubules. Thus, we determined the effects of stathmin individually phosphorylated at either Ser16 or Ser63 and doubly phosphorylated at both Ser25 and Ser38 on dynamic instability at plus and minus ends in vitro at microtubule polymer steady state and physiological pH (pH 7.2). We find that phosphorylation of Ser16 strongly reduces the direct catastrophe-promoting activity of stathmin at plus ends and abolishes it at minus ends, whereas phosphorylation of Ser63 abolishes the activity at both ends. The effects of phosphorylation of individual serines correlated well with stathmin''s reduced abilities to form stable T2S complexes, to inhibit microtubule polymerization, and to bind to microtubules. In contrast, double phosphorylation of Ser25 and Ser38 did not alter the ability of stathmin to modulate dynamic instability at the microtubule ends, its ability to form a stable T2S complex, or its ability to bind to microtubules. The data further support the hypotheses that phosphorylation of stathmin on either Ser16 or Ser63 plays a critical role in regulating microtubule polymerization and dynamics in cells.  相似文献   

5.
Kinase mediated phosphorylation signaling is extensively involved in cellular functions and human diseases, and unraveling phosphorylation networks requires the identification of substrates targeted by kinases, which has remained challenging. We report here a novel proteomic strategy to identify the specificity and direct substrates of kinases by coupling phosphoproteomics with a sensitive stable isotope labeled kinase reaction. A whole cell extract was moderately dephosphorylated and subjected to in vitro kinase reaction under the condition in which 18O-ATP is the phosphate donor. The phosphorylated proteins are then isolated and identified by mass spectrometry, in which the heavy phosphate (+85.979 Da) labeled phosphopeptides reveal the kinase specificity. The in vitro phosphorylated proteins with heavy phosphates are further overlapped with in vivo kinase-dependent phosphoproteins for the identification of direct substrates with high confidence. The strategy allowed us to identify 46 phosphorylation sites on 38 direct substrates of extracellular signal-regulated kinase 1, including multiple known substrates and novel substrates, highlighting the ability of this high throughput method for direct kinase substrate screening.Protein phosphorylation regulates almost all aspects of cell life, such as cell cycle, migration, and apoptosis (1), and deregulation of protein phosphorylation is one of the most frequent causes or consequences of human diseases including cancers, diabetes, and immune disorders (2). Up till now, however, known substrates are far from saturation for the majority of protein kinases (3); thus, mapping comprehensive kinase-substrate relationships is essential to understanding biological mechanisms and uncovering new drug targets (4).Accompanied with advances of high-speed and high-resolution mass spectrometry, the technique of kinase substrate screening using proteomic strategy is quickly evolving (57). Mass spectrometry has been extensively used for kinase-substrate interaction mapping (8) and global phosphorylation profiling (9). Although thousands of phosphorylation sites have been detected, complex phosphorylation cascade and crosstalk between pathways make it difficult for large-scale phosphoproteomics to reveal direct relationships between protein kinases and their substrates (10, 11). Extensive statistics, bioinformatics, and downstream biochemical assays are mandatory for the substrate verification (12, 13). Another strategy uses purified, active kinases to phosphorylate cell extracts in vitro, followed by mass spectrometric analysis to identify phosphoproteins. This approach inevitably faces the major challenge of separating real sites phosphorylated by target kinase and the phosphorylation triggered by endogenous kinases from cell lysates (14). Analog-sensitive kinase allele (15) overcomes the issue by utilizing the engineered kinase that can exclusively take a bulky-ATP analog under the reaction condition. Analog-sensitive kinase allele has been coupled with γ-thiophosphate analog ATP to facilitate the mass spectrometric analysis (1618).We have introduced kinase assay-linked phosphoproteomics (KALIP)1 to link the in vitro substrate identification and physiological phosphorylation events together in a high throughput manner (19, 20). The strategy, however, has only been applied to identify direct substrates of tyrosine kinases. In this study, we expanded the application of KALIP to serine/threonine kinases by introducing a quantitative strategy termed Stable Isotope Labeled Kinase Assay-Linked Phosphoproteomics (siKALIP). The method was applied to identify direct substrates of extracellular signal-regulated kinase 1 (ERK1), a serine/threonine kinase acting as an essential component of the Mitogen-activated protein kinase (MAPK) signal transduction pathway (21). A defect in the MAP/ERK pathway causes uncontrolled growth, which likely leads to cancer (22) and other diseases (2325). ERK1 can be activated by growth factors such as platelet-derived growth factor (PDGF), epidermal growth factor (EGF), and nerve growth factor (NGF) (26). Upon stimulation, ERK1 phosphorylates hundreds of substrates in various cellular compartments including cytoplasm, nucleus, and membrane (27). Among 38 ERK1 direct substrates identified by siKALIP, more than one third are previously discovered by classical molecular biology approaches, highlighting high specificity and sensitivity of the strategy. The results also support the hypothesis that ERK1 plays complex roles in multiple pathways that are essential for the cell growth regulation.  相似文献   

6.
7.
Reversible protein phosphorylation is a key regulatory mechanism of mitotic progression. Importantly, protein kinases themselves are also regulated by phosphorylation-dephosphorylation processes; hence, phosphorylation dynamics of kinases hold a wealth of information about phosphorylation networks. Here, we investigated the site-specific phosphorylation dynamics of human kinases during mitosis using synchronization of HeLa suspension cells, kinase enrichment, and high resolution mass spectrometry. In biological triplicate analyses, we identified 206 protein kinases and more than 900 protein kinase phosphorylation sites, including 61 phosphorylation sites on activation segments, and quantified their relative abundances across three specific mitotic stages. Around 25% of the kinase phosphorylation site ratios were found to be changed by at least 50% during mitotic progression. Further network analysis of jointly regulated kinase groups suggested that Cyclin-dependent kinase- and mitogen-activated kinase-centered interaction networks are coordinately down- and up-regulated in late mitosis, respectively. Importantly, our data cover most of the already known mitotic kinases and, moreover, identify attractive candidates for future studies of phosphorylation-based mitotic signaling. Thus, the results of this study provide a valuable resource for cell biologists and provide insight into the system properties of the mitotic phosphokinome.Reversible phosphorylation is a ubiquitous posttranslational protein modification that is involved in the regulation of almost all biological processes (13). In human, 518 protein kinases have been identified in the genome that phosphorylate the majority of cellular proteins and increase the diversity of the proteome by severalfold (4). Addition of a phosphate group to a protein can alter its structural, catalytic, and functional properties; hence, kinases require tight regulation to avoid unspecific phosphorylation, which can be deleterious to cells (57). As a result, cells use a variety of mechanisms to ensure proper regulation of kinase activities (8). Importantly, most kinases are also in turn regulated through autophosphorylation and phosphorylation by other kinases, thus generating complex phosphorylation networks. In particular, phosphorylation on activation segments is a common mechanism to modulate kinase activities (911), but additional phosphorylation sites are also frequently required for fine tuning of kinase localizations and functions (12). Some kinases contain phosphopeptide binding domains that recognize prephosphorylated sites on other kinases, resulting in processive phosphorylation and/or targeting of kinases to distinct cellular locations (1316). Because such priming phosphorylation events depend on the activities of the priming kinases, these motifs act as conditional docking sites and restrict the interaction with docking kinases to a particular point in time and physiological state. In addition, phosphorylation sites may act through combinatorial mechanisms or through cross-talk with other posttranslational modifications (PTMs)1 (17, 18), thus further increasing the complexity of kinase regulatory networks.Regulation of kinases is of particular interest in mitosis as most of the mitotic events are regulated by reversible protein phosphorylation (19). During mitosis, error-free segregation of sister chromatids into the two daughter cells is essential to ensure genomic stability. Physically, this process is carried out by the mitotic spindle, a highly dynamic microtubule-based structure. After entry into mitosis, the major microtubule-organizing centers in animal cells, the centrosomes, start to increase microtubule nucleation and move to opposite poles of the cell. Throughout prometaphase, microtubules emanating from centrosomes are captured by kinetochores, protein complexes assembled on centromeric chromosomal DNA. This eventually leads to the alignment of all chromosomes in a metaphase plate. Because proper bipolar attachment of chromosomes to spindle microtubules is essential for the correct segregation of chromosomes, this critical step is monitored by a signaling pathway known as the spindle assembly checkpoint (SAC) (20). This checkpoint is silenced only after all chromosomes have attached to the spindle in a bioriented fashion, resulting in the synchronous segregation of sister chromatids during anaphase. Simultaneously, a so-called central spindle is formed between the separating chromatids, and the formation of a contractile ring initiates cytokinesis. Finally, in telophase, the chromosomes decondense and reassemble into nuclei, whereas remnants of the central spindle form the midbody, marking the site of abscission. Cyclin-dependent kinase 1 (Cdk1), an evolutionarily conserved master mitotic kinase, is activated prior to mitosis and initiates most of the mitotic events. Cdk1 works in close association with other essential mitotic kinases such as Plk1, Aurora A, and Aurora B for the regulation of mitotic progression (19, 2124). Plk1 and Aurora kinases dynamically localize to different subcellular locations to perform multiple functions during mitosis and are phosphorylated at several conserved sites. Although little is known about the precise roles of these phosphorylation sites, emerging data indicate that they are involved in regulating localization-specific functions (25, 26). Furthermore, the kinases Bub1, BubR1, and TTK (Mps1) and kinases of the Nek family play important roles in maintaining the fidelity and robustness of mitosis (19). Recently, a genome-wide RNA-mediated interference screen identified M phase phenotypes for many kinases that have not previously been implicated in cell cycle functions, indicating that additional kinases have important mitotic functions (27).Although protein phosphorylation plays a pivotal role in the regulation of cellular networks, many phosphorylation events remain undiscovered mainly because of technical limitations (28). The advent of mass spectrometry-based proteomics along with developments in phosphopeptide enrichment methods has enabled large scale global phosphoproteomics studies (29, 30). However, the number of phosphorylation sites identified on kinases is limited compared with other proteins because of their frequently low expression levels. To overcome this problem, small inhibitor-based kinase enrichment strategies were developed, resulting in the identification of more than 200 kinases from HeLa cell lysates (31, 32). This method was also used recently to compare the phosphokinomes during S phase and M phase of the cell cycle, resulting in the identification of several hundreds of M phase-specific kinase phosphorylation sites (31). In the present study, we address the dynamics of the phosphokinome during mitotic progression using large scale cell synchronization at three distinct mitotic stages, small inhibitor-based kinase enrichment, and stable isotope labeling by amino acids in cell culture (SILAC)-based quantitative mass spectrometry. Thus, we determined the mitotic phosphorylation dynamics of more than 900 kinase phosphorylation sites and identified distinctly regulated kinase interaction networks. Our results provide a valuable resource for the dynamics of the kinome during mitotic progression and give insight into the system properties of kinase interaction networks.  相似文献   

8.
9.
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   

10.
11.
12.
Relationships among biochemical signaling processes involved in Ca2+/calmodulin (CaM)-dependent phosphorylation of smooth muscle myosin regulatory light chain (RLC) by myosin light chain kinase (MLCK) were determined. A genetically-encoded biosensor MLCK for measuring Ca2+-dependent CaM binding and activation was expressed in smooth muscles of transgenic mice. We performed real-time evaluations of the relationships among [Ca2+]i, MLCK activation, and contraction in urinary bladder smooth muscle strips neurally stimulated for 3 s. Latencies for the onset of [Ca2+]i and kinase activation were 55 ± 8 and 65 ± 6 ms, respectively. Both increased with RLC phosphorylation at 100 ms, whereas force latency was 109 ± 3 ms. [Ca2+]i, kinase activation, and RLC phosphorylation responses were maximal by 1.2 s, whereas force increased more slowly to a maximal value at 3 s. A delayed temporal response between RLC phosphorylation and force is probably due to mechanical effects associated with elastic elements in the tissue. MLCK activation partially declined at 3 s of stimulation with no change in [Ca2+]i and also declined more rapidly than [Ca2+]i during relaxation. The apparent desensitization of MLCK to Ca2+ activation appears to be due to phosphorylation in its calmodulin binding segment. Phosphorylation of two myosin light chain phosphatase regulatory proteins (MYPT1 and CPI-17) or a protein implicated in strengthening membrane adhesion complexes for force transmission (paxillin) did not change during force development. Thus, neural stimulation leads to rapid increases in [Ca2+]i, MLCK activation, and RLC phosphorylation in phasic smooth muscle, showing a tightly coupled Ca2+ signaling complex as an elementary mechanism initiating contraction.Increases in [Ca2+]i3 in smooth muscle cells lead to Ca2+/CaM-dependent MLCK activation and RLC phosphorylation. Phosphorylation of RLC increases actin-activated myosin MgATPase activity leading to myosin cross-bridge cycling with force development (13).The activation of smooth muscle contraction may be affected by multiple cellular processes. Previous investigations show that free Ca2+/CaM is limiting for kinase activation despite the abundance of total CaM (46). The extent of RLC phosphorylation is balanced by the actions of MLCK and myosin light chain phosphatase, which is composed of three distinct protein subunits (7). The myosin phosphatase targeting subunit, MYPT1, in smooth muscle binds to myosin filaments, thus targeting the 37-kDa catalytic subunit (type 1 serine/threonine phosphatase, PP1c) to phosphorylated RLC. RLC phosphorylation and muscle force may be regulated by additional signaling pathways involving phosphorylation of RLC by Ca2+-independent kinase(s) and inhibition of myosin light chain phosphatase, processes that increase the contraction response at fixed [Ca2+]i (Ca2+-sensitization) (814). Many studies indicate that agonist-mediated Ca2+-sensitization most often reflects decreased myosin light chain phosphatase activity involving two major pathways including MYPT1 phosphorylation by a Rho kinase pathway and phosphorylation of CPI-17 by PKC (8, 1416). Additionally, phosphorylation of MLCK in its calmodulin-binding sequence by a Ca2+/calmodulin-dependent kinase pathway has been implicated in Ca2+ desensitization of RLC phosphorylation (1719). How these signaling pathways intersect the responses of the primary Ca2+/CaM pathway during physiological neural stimulation is not known.There is also evidence that smooth muscle contraction requires the polymerization of submembranous cytoskeletal actin filaments to strengthen membrane adhesion complexes involved in transmitting force between actin-myosin filaments and external force-transmitting structures (2023). In tracheal smooth muscle, paxillin at membrane adhesions undergoes tyrosine phosphorylation in response to contractile stimulation by an agonist, and this phosphorylation increases concurrently with force development in response to agonist. Expression of nonphosphorylatable paxillin mutants in tracheal muscle suppresses acetylcholine-induced tyrosine phosphorylation of paxillin, tension development, and actin polymerization without affecting RLC phosphorylation (24, 25). Thus, paxillin phosphorylation may play an important role in tension development in smooth muscle independently of RLC phosphorylation and cross-bridge cycling.Specific models relating signaling mechanisms in the smooth muscle cell to contraction dynamics are limited when cells in tissues are stimulated slowly and asynchronously by agonist diffusing into the preparation. Field stimulation leading to the rapid release of neurotransmitters from nerves embedded in the tissue avoids these problems associated with agonist diffusion (26, 27). In urinary bladder smooth muscle, phasic contractions are brought about by the parasympathetic nervous system. Upon activation, parasympathetic nerve varicosities release the two neurotransmitters, acetylcholine and ATP, that bind to muscarinic and purinergic receptors, respectively. They cause smooth muscle contraction by inducing Ca2+ transients as elementary signals in the process of nerve-smooth muscle communication (2830). We recently reported the development of a genetically encoded, CaM-sensor for activation of MLCK. The CaM-sensor MLCK contains short smooth muscle MLCK fused to two fluorophores, enhanced cyan fluorescent protein and enhanced yellow fluorescent protein, linked by the MLCK calmodulin binding sequence (6, 14, 31). Upon dimerization, there is significant FRET from the donor enhanced cyan fluorescent protein to the acceptor enhanced yellow fluorescent protein. Ca2+/CaM binding dissociates the dimer resulting in a decrease in FRET intensity coincident with activation of kinase activity (31). Thus, CaM-sensor MLCK is capable of directly monitoring Ca2+/CaM binding and activation of the kinase in smooth muscle tissues where it is expressed specifically in smooth muscle cells of transgenic mice. We therefore combined neural stimulation with real-time measurements of [Ca2+]i, MLCK activation, and force development in smooth muscle tissue from these mice. Additionally, RLC phosphorylation was measured precisely at specific times following neural stimulation in tissues frozen by a rapid-release electronic freezing device (26, 27). Results from these studies reveal that physiological stimulation of smooth muscle cells by neurotransmitter release leads to rapid increases in [Ca2+]i, MLCK activation, and RLC phosphorylation at similar rates without the apparent activities of Ca2+-independent kinases, inhibition of myosin light chain phosphatase, or paxillin phosphorylation. Thus, the elemental processes for phasic smooth muscle contraction are represented by this tightly coupled Ca2+ signaling complex.  相似文献   

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.
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   

15.
The reversible regulation of myosin light chain phosphatase (MLCP) in response to agonist stimulation and cAMP/cGMP signals plays an important role in the regulation of smooth muscle (SM) tone. Here, we investigated the mechanism underlying the inhibition of MLCP induced by the phosphorylation of myosin phosphatase targeting subunit (MYPT1), a regulatory subunit of MLCP, at Thr-696 and Thr-853 using glutathione S-transferase (GST)-MYPT1 fragments having the inhibitory phosphorylation sites. GST-MYPT1 fragments, including only Thr-696 and only Thr-853, inhibited purified MLCP (IC50 = 1.6 and 60 nm, respectively) when they were phosphorylated with RhoA-dependent kinase (ROCK). The activities of isolated catalytic subunits of type 1 and type 2A phosphatases (PP1 and PP2A) were insensitive to either fragment. Phospho-GST-MYPT1 fragments docked directly at the active site of MLCP, and this was blocked by a PP1/PP2A inhibitor microcystin (MC)-LR or by mutation of the active sites in PP1. GST-MYPT1 fragments induced a contraction of β-escin-permeabilized ileum SM at constant pCa 6.3 (EC50 = 2 μm), which was eliminated by Ala substitution of the fragment at Thr-696 or by ROCK inhibitors or 8Br-cGMP. GST-MYPT1-(697–880) was 5-times less potent than fragments including Thr-696. Relaxation induced by 8Br-cGMP was not affected by Ala substitution at Ser-695, a known phosphorylation site for protein kinase A/G. Thus, GST-MYPT1 fragments are phosphorylated by ROCK in permeabilized SM and mimic agonist-induced inhibition and cGMP-induced activation of MLCP. We propose a model in which MYPT1 phosphorylation at Thr-696 and Thr-853 causes an autoinhibition of MLCP that accounts for Ca2+ sensitization of smooth muscle force.The contractile state of smooth muscle (SM)3 is driven by phosphorylation of the regulatory myosin light chain and reflects the balance of the Ca2+-calmodulin-dependent myosin light chain kinase and myosin light chain phosphatase (MLCP) activities (1). The stoichiometry between force and [Ca2+] varies with different agonists (2), reflecting other signaling pathways that modulate the MLCP or myosin light chain kinase activities (35). Agonist activation of G-protein-coupled receptors triggers Ca2+ release from the sarcoplasmic reticulum. Simultaneously, G-protein-coupled receptor signals are mediated by Ca2+-independent phospholipase A2 (6) and initiate kinase signals, such as PKC, phosphoinositide 3-kinase (7), and ROCK. These lead to inhibition of MLCP activity resulting in an increase in regulatory myosin light chain phosphorylation independent of a change in Ca2+ (Ca2+ sensitization) (for review, see Ref. 1). K+ depolarization can also activate RhoA in a Ca2+-dependent manner (8). Conversely, Ca2+ desensitization occurs when nitric oxide production and the activation of Gas elevate cGMP and cAMP levels in SM, leading to dis-inhibition and restoration of MLCP activity (915). Thus, MLCP plays a pivotal role in controlling phosphorylation of myosin, in response to physiological stimulation.MLCP is a trimeric holoenzyme consisting of a catalytic subunit of protein phosphatase 1 (PP1) δ isoform and a regulatory complex of MYPT1 and an accessory M21 subunit (16). A PP1 binding site, KVKF38, is located at the N terminus of MYPT1 followed by an ankyrin-repeat domain. This N-terminal domain forms a part of the active site together with the catalytic subunit and controls the substrate specificity via allosteric interaction and targeting to loci (17). The C-terminal region of MYPT1 directly binds to substrates such as myosin and ezrin/radixin/moecin proteins as well as, under some conditions, the plasma membrane, tethering the catalytic subunit to multiple targets (18, 19). Furthermore, MYPT1 is involved in the regulation of MLCP activity. Alternative splicing of MYPT1 occurs in SM depending on the tissue and the developmental stage (20). An exon 13 splicing of MYPT1 is involved in Ca2+ sensitization that occurs in response to GTP (21), whereas a splice variant of MYPT1, containing the C-terminal Leu-zipper sequence, correlates with cGMP-dependent relaxation of smooth muscle (22). Direct binding of PKG to MYPT1 at the Leu-zipper domain and/or Arg/Lys-rich domain is involved in the activation of MLCP (2325). In addition, a myosin phosphatase-Rho interacting protein (M-RIP) is directly associated with the MYPT1 C-terminal domain, proposed to recruit RhoA to the MLCP complex (26). The C-terminal region also binds to ZIP kinase, which phosphorylates MYPT1 at Thr-6964 (27). Thus, the C-terminal domain of MYPT1 functions as a scaffold for multiple phosphatase regulatory proteins.Phosphorylation of MYPT1 at Thr-696 and Thr-853 and the phosphatase inhibitory protein CPI-17 at Thr-38 play dominant roles in the agonist-induced inhibition of MLCP (18, 2834), yet the molecular mechanism(s) of MYPT1 inhibitory phosphorylation is poorly understood. Receptor activation induces biphasic contraction of SM, reflecting a sequential activation of PKC and ROCK. Phosphorylation of CPI-17 occurs first in parallel with Ca2+ release and the activation of a conventional PKC that causes Ca2+-dependent Ca2+ sensitization (35). A delayed activation of ROCK increases the phosphorylation of MYPT1 at Thr-853. These phosphorylation events maintain the sustained phase of contraction after the fall in [Ca2+]i (35). Phosphorylation of MYPT1 at Thr-853 is elevated in response to various agonists (35, 36). Unlike the Thr-853 site, phosphorylation of MYPT1 at Thr-696 is often spontaneously phosphorylated under resting conditions and insensitive to stimuli with most agonists (36). Nonetheless, up-regulation of MYPT1 phosphorylation at Thr-696 is reported in some types of hypertensive animals and patients, suggesting an importance of the site under pathological conditions (3739). Phosphorylation of CPI-17 and MYPT1 at Thr-696 is reversed in response to nitric oxide production and cGMP elevation, which parallels relaxation (14, 15). Upon cGMP elevation, MYPT1 at Ser-695 is phosphorylated, and the Ser phosphorylation blocks the adjacent phosphorylation at Thr-696, causing dis-inhibition of MLCP (27, 40). However, Ser-695 phosphorylation does not cause the dephosphorylation at Thr-696 in intact cerebral artery (41). Thus, phosphorylation of MYPT1 governs Ca2+ sensitization and desensitization of SM, although the underlying mechanisms are still controversial. In addition, telokin, a dominant protein in visceral and phasic vascular SM tissues, is phosphorylated by PKG and PKA, activating MLCP by an unknown mechanism and inducing SM relaxation (42).Multiple mechanisms have been suggested for the phosphorylation-dependent inhibition of MLCP. Thiophosphorylation of MYPT1 results in lower Vm and higher Km values of MLCP activity, suggesting that allosteric modulation of the active site is necessary for the thiophosphorylation-dependent inhibition of MLCP (43). On the other hand, translocation of MYPT1 to the plasma membrane region occurs in parallel with the phosphorylation of MYPT1 at Thr-696 (44, 45), but the amount translocated and the functional meaning remain controversial (41). Phosphorylation of MYPT1 at Thr-853 in vitro reduces its affinity for phospho-myosin, thus suppressing the phosphatase activity (18). It has also been demonstrated that reconstitution of thiophosphorylated MYPT1 at Thr-696 or Thr-853 with isolated PP1δ produces a less-active form of MLCP complex (46). This supports the kinetic analysis (43) that suggests an allosteric effect of MYPT1 phosphorylation on the phosphatase activity. In contrast, a thiophosphopeptide mimicking the phosphorylation site of MBS85, a homolog of MYPT1 and not present in SM, inhibits the activity of MBS85·PP1 complex, suggesting the direct interaction between the MBS85 site and PP1 (47). In the crystal structure model of MYPT1-(1–229). PP1δ complex, the electrostatic potential map at the MLCP active site complements amino acid profiles around the phosphorylation sites (17). Therefore, it is possible that the inhibitory phosphorylation sites directly dock at the active site of MLCP and inhibit the activity. Here, we examine mechanisms underlying the inhibition of MLCP through the phosphorylation of MYPT1 at Thr-696 and Thr-853 using GST fusion versions of various MYPT1 fragments including or excluding either or both of these phosphorylation sites. Phosphorylated MYPT1 fragments including either Thr-696 or Thr-853 potently and specifically inhibit MLCP purified from pig aorta and the enzyme associated with myofilaments in permeabilized ileum SM tissues. We further show that inhibition of MLCP in SM tissues is eliminated by activation of PKA/PKG, suggesting that the GST-MYPT1 fragments mimic agonist-induced autoinhibition and cAMP/cGMP-dependent dis-autoinhibition of MLCP in SM.  相似文献   

16.
17.
18.
19.
The Dbf4-Cdc7 kinase (DDK) is required for the activation of the origins of replication, and DDK phosphorylates Mcm2 in vitro. We find that budding yeast Cdc7 alone exists in solution as a weakly active multimer. Dbf4 forms a likely heterodimer with Cdc7, and this species phosphorylates Mcm2 with substantially higher specific activity. Dbf4 alone binds tightly to Mcm2, whereas Cdc7 alone binds weakly to Mcm2, suggesting that Dbf4 recruits Cdc7 to phosphorylate Mcm2. DDK phosphorylates two serine residues of Mcm2 near the N terminus of the protein, Ser-164 and Ser-170. Expression of mcm2-S170A is lethal to yeast cells that lack endogenous MCM2 (mcm2Δ); however, this lethality is rescued in cells harboring the DDK bypass mutant mcm5-bob1. We conclude that DDK phosphorylation of Mcm2 is required for cell growth.The Cdc7 protein kinase is required throughout the yeast S phase to activate origins (1, 2). The S phase cyclin-dependent kinase also activates yeast origins of replication (35). It has been proposed that Dbf4 activates Cdc7 kinase in S phase, and that Dbf4 interaction with Cdc7 is essential for Cdc7 kinase activity (6). However, it is not known how Dbf4-Cdc7 (DDK)2 acts during S phase to trigger the initiation of DNA replication. DDK has homologs in other eukaryotic species, and the role of Cdc7 in activation of replication origins during S phase may be conserved (710).The Mcm2-7 complex functions with Cdc45 and GINS to unwind DNA at a replication fork (1115). A mutation of MCM5 (mcm5-bob1) bypasses the cellular requirements for DBF4 and CDC7 (16), suggesting a critical physiologic interaction between Dbf4-Cdc7 and Mcm proteins. DDK phosphorylates Mcm2 in vitro with proteins purified from budding yeast (17, 18) or human cells (19). Furthermore, there are mutants of MCM2 that show synthetic lethality with DBF4 mutants (6, 17), suggesting a biologically relevant interaction between DBF4 and MCM2. Nevertheless, the physiologic role of DDK phosphorylation of Mcm2 is a matter of dispute. In human cells, replacement of MCM2 DDK-phosphoacceptor residues with alanines inhibits DNA replication, suggesting that Dbf4-Cdc7 phosphorylation of Mcm2 in humans is important for DNA replication (20). In contrast, mutation of putative DDK phosphorylation sites at the N terminus of Schizosaccharomyces pombe Mcm2 results in viable cells, suggesting that phosphorylation of S. pombe Mcm2 by DDK is not critical for cell growth (10).In budding yeast, Cdc7 is present at high levels in G1 and S phase, whereas Dbf4 levels peak in S phase (18, 21, 22). Furthermore, budding yeast DDK binds to chromatin during S phase (6), and it has been shown that Dbf4 is required for Cdc7 binding to chromatin in budding yeast (23, 24), fission yeast (25), and Xenopus (9). Human and fission yeast Cdc7 are inert on their own (7, 8), but Dbf4-Cdc7 is active in phosphorylating Mcm proteins in budding yeast (6, 26), fission yeast (7), and human (8, 10). Based on these data, it has been proposed that Dbf4 activates Cdc7 kinase in S phase and that Dbf4 interaction with Cdc7 is essential for Cdc7 kinase activity (6, 9, 18, 2124). However, a mechanistic analysis of how Dbf4 activates Cdc7 has not yet been accomplished. For example, the multimeric state of the active Dbf4-Cdc7 complex is currently disputed. A heterodimer of fission yeast Cdc7 (Hsk1) in complex with fission yeast Dbf4 (Dfp1) can phosphorylate Mcm2 (7). However, in budding yeast, oligomers of Cdc7 exist in the cell (27), and Dbf4-Cdc7 exists as oligomers of 180 and 300 kDa (27).DDK phosphorylates the N termini of human Mcm2 (19, 20, 28), human Mcm4 (10), budding yeast Mcm4 (26), and fission yeast Mcm6 (10). Although the sequences of the Mcm N termini are poorly conserved, the DDK sites identified in each study have neighboring acidic residues. The residues of budding yeast Mcm2 that are phosphorylated by DDK have not yet been identified.In this study, we find that budding yeast Cdc7 is weakly active as a multimer in phosphorylating Mcm2. However, a low molecular weight form of Dbf4-Cdc7, likely a heterodimer, has a higher specific activity for phosphorylation of Mcm2. Dbf4 or DDK, but not Cdc7, binds tightly to Mcm2, suggesting that Dbf4 recruits Cdc7 to Mcm2. DDK phosphorylates two serine residues of Mcm2, Ser-164 and Ser-170, in an acidic region of the protein. Mutation of Ser-170 is lethal to yeast cells, but this phenotype is rescued by the DDK bypass mutant mcm5-bob1. We conclude that DDK phosphorylation of Ser-170 of Mcm2 is required for budding yeast growth.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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