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 15 毫秒
1.
2.
3.
We expressed the γ-subspecies of protein kinase C (γ-PKC) fused with green fluorescent protein (GFP) in various cell lines and observed the movement of this fusion protein in living cells under a confocal laser scanning fluorescent microscope. γ-PKC–GFP fusion protein had enzymological properties very similar to that of native γ-PKC. The fluorescence of γ-PKC– GFP was observed throughout the cytoplasm in transiently transfected COS-7 cells. Stimulation by an active phorbol ester (12-O-tetradecanoylphorbol 13-acetate [TPA]) but not by an inactive phorbol ester (4α-phorbol 12, 13-didecanoate) induced a significant translocation of γ-PKC–GFP from cytoplasm to the plasma membrane. A23187, a Ca2+ ionophore, induced a more rapid translocation of γ-PKC–GFP than TPA. The A23187-induced translocation was abolished by elimination of extracellular and intracellular Ca2+. TPA- induced translocation of γ-PKC–GFP was unidirected, while Ca2+ ionophore–induced translocation was reversible; that is, γ-PKC–GFP translocated to the membrane returned to the cytosol and finally accumulated as patchy dots on the plasma membrane. To investigate the significance of C1 and C2 domains of γ-PKC in translocation, we expressed mutant γ-PKC–GFP fusion protein in which the two cysteine rich regions in the C1 region were disrupted (designated as BS 238) or the C2 region was deleted (BS 239). BS 238 mutant was translocated by Ca2+ ionophore but not by TPA. In contrast, BS 239 mutant was translocated by TPA but not by Ca2+ ionophore. To examine the translocation of γ-PKC–GFP under physiological conditions, we expressed it in NG-108 cells, N-methyl-d-aspartate (NMDA) receptor–transfected COS-7 cells, or CHO cells expressing metabotropic glutamate receptor 1 (CHO/mGluR1 cells). In NG-108 cells , K+ depolarization induced rapid translocation of γ-PKC–GFP. In NMDA receptor–transfected COS-7 cells, application of NMDA plus glycine also translocated γ-PKC–GFP. Furthermore, rapid translocation and sequential retranslocation of γ-PKC–GFP were observed in CHO/ mGluR1 cells on stimulation with the receptor. Neither cytochalasin D nor colchicine affected the translocation of γ-PKC–GFP, indicating that translocation of γ-PKC was independent of actin and microtubule. γ-PKC–GFP fusion protein is a useful tool for investigating the molecular mechanism of γ-PKC translocation and the role of γ-PKC in the central nervous system.Protein kinase C (PKC),1 a family of phospholipid-dependent serine/threonine kinases and of which there are at least 12 subspecies, plays an important role in various cellular signal transductions (Nishizuka, 1984, 1988, 1992). Regardless of ubiquitous expression of PKCs in various tissues, the central nervous system abundantly contains several unique PKCs. In particular, the γ-subspecies of PKC (γ-PKC) is present only in the central nervous system and is thought to be involved in many neuronal functions including the formation of neural plasticity and memory (Nishizuka, 1986; Abeliovich et al., 1993a ,b; Tanaka and Nishizuka, 1994).PKC isozymes are divided into three subfamilies based on differences in the regulatory domain: conventional PKC (cPKC), novel PKC (nPKC), and atypical PKC (aPKC). Conventional PKCs have two common regions in the regulatory domain, C1 and C2. The C1 region has two cysteine-rich loops (zinc finger–like motifs) that interact with diacylglycerol (DG) or phorbol esters (Nishizuka, 1988; Ono et al., 1989). The C2 region mediates calcium binding (Ono et al., 1989) and is only present in cPKCs (Ono et al., 1988b ), although a region related to C2 region was recently reported in nPKC, a calcium-independent PKC (Parker and Dekker, 1997). Full activation of cPKCs, including γ-PKC, requires DG and calcium. The C1 region is also present in nPKC, and one of the cysteine-rich loops is found in aPKCs.Conventional PKCs and nPKCs, whose regulatory domains contain C1, are known to be translocated from the cytosol to particulate fraction when activated by DG or phorbol esters (Kraft et al., 1982). Therefore, the translocation of PKCs is a good marker of whether these enzymes are activated. Although this phenomenon is well known, the mechanism and physiological significance of PKC translocation have not yet been clarified. By conventional enzymological or immunohistochemical methods, it is impossible to observe the translocation of PKC in real time, in the same cells, and in living states, except in the investigation using fluorescent probes that directly bind PKC (Chen and Poenie, 1993). In addition, these fluorescent compounds are suggested to inhibit the activity of PKC itself at high concentration.To resolve these problems and to directly observe the translocation of γ-PKC in living cells, we produced a fusion protein of γ-PKC and green fluorescent protein (GFP). The GFP, isolated from jellyfish Aequorea victoria, has fluorescence without additional substrates and cofactors (Cubitt et al., 1995). Recent studies have revealed that GFP is a good candidate as a molecular reporter protein to monitor the alternation of protein localization, gene expression, and protein trafficking in living cells (Cubitt et al., 1995). In this study, we visualized and analyzed the translocation of γ-PKC–GFP fusion protein with confocal laser scanning fluorescence microscopy, using various stimulations, such as phorbol esters, Ca2+ ionophore, K+ depolarization, and receptor-mediated stimulus.  相似文献   

4.
Integrin signaling plays critical roles in cell adhesion, spreading, and migration, and it is generally accepted that to regulate these integrin functions accurately, localized actin remodeling is required. However, the molecular mechanisms that control the targeting of actin regulation molecules to the proper sites are unknown. We previously demonstrated that integrin-mediated cell spreading and migration on fibronectin are dependent on the localized activation of phospholipase D (PLD). However, the mechanism underlying PLD activation by integrin is largely unknown. Here we demonstrate that protein kinase Cδ (PKCδ) is required for integrin-mediated PLD signaling. After integrin stimulation, PKCδ is activated and translocated to the edges of lamellipodia, where it colocalizes with PLD2. The abrogation of PKCδ activity inhibited integrin-induced PLD activation and cell spreading. Finally, we show that Thr566 of PLD2 is directly phosphorylated by PKCδ and that PLD2 mutation in this region prevents PLD2 activation, PLD2 translocation to the edge of lamellipodia, Rac translocation, and cell spreading after integrin activation. Together, these results suggest that PKCδ is a primary regulator of integrin-mediated PLD activation via the direct phosphorylation of PLD, which is essential for directing integrin-induced cell spreading.Integrin-mediated cell adhesion, spreading, and migration, which are essential for cellular differentiation, proliferation, survival, chemotaxis, and wound healing, require cell polarization with an environmental stimulus (32). To regulate these integrin-mediated functions accurately, coordinated and spatial control of localized cytoskeletal rearrangement is required. The key downstream signaling molecules of integrin-mediated actin cytoskeletal rearrangements include small GTPases of the Rho family, such as Rho, Cdc42, and Rac (57, 58). Recently it was suggested that integrin indirectly regulates the recruitment of small G proteins and their localized activation at a specific plasma membrane region called the cholesterol-enriched membrane microdomain. Furthermore, the membrane targeting of these molecules appears to be required for the activation of downstream effectors that induce actin reorganization (8, 9, 48). However, in the absence of integrin signaling, despite the GTP loading status, activated Rac and Cdc42 remain in the cytosol and cannot activate downstream effectors, such as p21-activated kinase (PAK) (8). This regulation of the localization of small GTPases to a specific site is supported by the observation that Rac1 is localized and activated at the leading edges of migrating cells, while Cdc42 is also activated in cellular protrusions and in the peripheral region (33, 51). The differentially localized activation of small GTPases results in coordinated spatially confined signaling leading to cytoskeletal rearrangement, which is critical for the regulation of integrin-mediated cell spreading and directional cell migration.The hydrolysis of phosphatidylcholine by phospholipase D1 (PLD1) and PLD2 generates the messenger lipid phosphatidic acid (PA) in response to a variety of signals, which include hormones, neurotransmitters, and growth factors (17). It has been shown that PA affects actin cytoskeletal rearrangement and hence lamellipodium extension and integrin-mediated cell spreading as well as migration. PLD activity has been found in detergent-insoluble membrane fractions in which a wide variety of cytoskeletal proteins, such as F-actin, α-actinin, vinculin, paxillin, and talin, were enriched (34). Furthermore, the stimulation of PLD with physiologic and pharmacologic agonists results in its association with actin filaments (34). In addition, actin polymerization and stress fiber formation are tightly coupled to the activation of PLD (14). The formation of lamellipodium structures and membrane ruffles is blocked by PLD inhibition (53, 60), and PLD activity is critical for epithelial cell, leukocyte, and neutrophil adhesion and migration (41, 43, 52). Furthermore, we have previously shown that the activity of PLD is upregulated, and that the activated PLD is translocated to lamellipodia, after integrin activation (3). The PLD product PA acts as a lipid anchor for the membrane translocation of Rac, and this PA-mediated localized activation of Rac is critical for integrin-mediated cell spreading and migration through Rac downstream signaling activation and actin cytoskeleton rearrangement (3). However, the mechanisms that regulate the activation and localization of PLD, which induce the localized downstream activation of integrin signaling, have not been elucidated.Members of the protein kinase C (PKC) family of serine-threonine kinases are known to play important roles in the transduction of signals from the activation of integrin to cell adhesion and spreading, as well as in cell migration via actin reorganization (11, 25, 61, 66). Several studies have shown that the activities of several PKC isozymes are modulated and are crucially required for integrin-mediated cell spreading and migration. The PKCα, -δ, and -ɛ isotypes were activated and then translocated from the cytosol to the membrane after integrin activation, and inhibition of these PKC isozymes prevented cell spreading (10, 47, 66). In addition, the activation of PKCα, -δ, and -ɛ rescued the spreading of α5 integrin-deficient cells on fibronectin (10), and PKCβΙ mediated platelet cell spreading and migration on fibrinogen (2). It has also been demonstrated that PKCθ activity is involved in endothelial cell migration (65). These results suggest that the kinase activities of diverse members of PKC are involved in the integrin-mediated signaling pathway leading to the actin cytoskeletal rearrangement required for cell spreading and migration. Several PKC substrates are known to influence the actin cytoskeleton directly (42). However, the natures of the isoform-specific functions of PKC members and of their specific downstream effectors for actin cytoskeletal rearrangement induction by integrin signaling remain to be elucidated.In this study, we found that PKCδ is an upstream modulator of localized PLD activation in the integrin signaling pathway. We demonstrate for the first time that PKCδ activity (not PKCα or PKCɛ activity) is critical for integrin-mediated PLD activation, and we found that PLD2 is phosphorylated at Thr566 by PKCδ in the integrin signaling pathway. Furthermore, we show that this phosphorylation is critical for integrin-mediated targeting of PLD to membrane ruffles, Rac translocation to the membrane, and lamellipodium formation during cell spreading. These findings strongly suggest a bridge between PKCδ and the signaling of actin cytoskeletal rearrangement by the integrin signaling pathway via PLD activation, and they provide a novel molecular mechanism for localized PLD activation via PKCδ phosphorylation, which is critical for the actin cytoskeletal rearrangements required for integrin-mediated cell spreading.  相似文献   

5.
6.

Background

One of the crucial steps toward understanding the biological functions of a cellular system is to investigate protein–protein interaction (PPI) networks. As an increasing number of reliable PPIs become available, there is a growing need for discovering PPIs to reconstruct PPI networks of interesting organisms. Some interolog-based methods and homologous PPI families have been proposed for predicting PPIs from the known PPIs of source organisms.

Results

Here, we propose a multiple-strategy scoring method to identify reliable PPIs for reconstructing the mouse PPI network from two well-known organisms: human and fly. We firstly identified the PPI candidates of target organisms based on homologous PPIs, sharing significant sequence similarities (joint E-value ≤ 1 × 10−40), from source organisms using generalized interolog mapping. These PPI candidates were evaluated by our multiple-strategy scoring method, combining sequence similarities, normalized ranks, and conservation scores across multiple organisms. According to 106,825 PPI candidates in yeast derived from human and fly, our scoring method can achieve high prediction accuracy and outperform generalized interolog mapping. Experiment results show that our multiple-strategy score can avoid the influence of the protein family size and length to significantly improve PPI prediction accuracy and reflect the biological functions. In addition, the top-ranked and conserved PPIs are often orthologous/essential interactions and share the functional similarity. Based on these reliable predicted PPIs, we reconstructed a comprehensive mouse PPI network, which is a scale-free network and can reflect the biological functions and high connectivity of 292 KEGG modules, including 216 pathways and 76 structural complexes.

Conclusions

Experimental results show that our scoring method can improve the predicting accuracy based on the normalized rank and evolutionary conservation from multiple organisms. Our predicted PPIs share similar biological processes and cellular components, and the reconstructed genome-wide PPI network can reflect network topology and modularity. We believe that our method is useful for inferring reliable PPIs and reconstructing a comprehensive PPI network of an interesting organism.  相似文献   

7.
8.
Protein kinase Cδ (PKCδ) deficiency causes autoimmune pathology in humans and mice and is crucial for the maintenance of B cell homeostasis. However, the mechanisms underlying autoimmune disease in PKCδ deficiency remain poorly defined. Here, we address the antigen-dependent and -independent roles of PKCδ in B cell development, repertoire selection, and antigen responsiveness. We demonstrate that PKCδ is rapidly phosphorylated downstream of both the B cell receptor (BCR) and the B cell-activating factor (BAFF) receptor. We found that PKCδ is essential for antigen-dependent negative selection of splenic transitional B cells and is required for activation of the proapoptotic Ca2+-Erk pathway that is selectively activated during B cell-negative selection. Unexpectedly, we also identified a previously unrecognized role for PKCδ as a proximal negative regulator of BCR signaling that substantially impacts survival and proliferation of mature follicular B cells. As a consequence of these distinct roles, PKCδ deficiency leads to the survival and development of a B cell repertoire that is not only aberrantly autoreactive but also hyperresponsive to antigen stimulation.  相似文献   

9.
In this study, we used gene targeting in mice to identify the in vivo functions of PKD1. In addition to phenotypically characterizing the resulting knock-out animals, we also used mouse embryonic fibroblasts to investigate the associated signaling pathways in detail. This study is the first to use genetic deletion to reveal that PKD1 is a key regulator involved in determining the threshold of mitochondrial depolarization that leads to the production of reactive oxygen species. In addition, we also provide clear evidence that PKCδ is upstream of PKD1 in this process and acts as the activating kinase of PKD1. Therefore, our in vivo data indicate that PKD1 functions not only in the context of aging but also during nutrient deprivation, which occurs during specific phases of tumor growth.  相似文献   

10.
11.
The subcellular localization of protein kinase C (PKC)-δ was determined in HL60 cells differentiated toward monocytes/macrophages by treatment with TPA. PKC-δ was detected in the nucleus and cytoplasm of differentiated HL60 cells and, more specifically, associated with structures resembling intermediate filaments. Indirect immunostaining revealed that PKC-δ colocalized with vimentin in the cytosol and perinuclear region of these cells. Immunoprecipitation studies showed that PKC-δ was in an active (autophosphorylated) state in differentiated HL60 cells and that vimentin immunoprecipitated from these cells was also phosphorylated. Treatment of HL60 cells with the PKC-specific inhibitor chelerythrine decreased the phosphorylation of vimentin. These data suggest that vimentin is a substrate for PKC-δ and that this PKC isoenzyme may play a specific role in the regulation of shape change and cell adhesion during HL60 differentiation.  相似文献   

12.
Protein interaction networks comprise thousands of individual binary links between distinct proteins. Whilst these data have attracted considerable attention and been the focus of many different studies, the networks, their structure, function, and how they change over time are still not fully known. More importantly, there is still considerable uncertainty regarding their size, and the quality of the available data continues to be questioned. Here, we employ statistical models of the experimental sampling process, in particular capture–recapture methods, in order to assess the false discovery rate and size of protein interaction networks. We uses these methods to gauge the ability of different experimental systems to find the true binary interactome. Our model allows us to obtain estimates for the size and false-discovery rate from simple considerations regarding the number of repeatedly interactions, and provides suggestions as to how we can exploit this information in order to reduce the effects of noise in such data. In particular our approach does not require a reference dataset. We estimate that approximately more than half of the true physical interactome has now been sampled in yeast.  相似文献   

13.
To better study the role of PKCδ in normal function and disease, we developed an ATP analog-specific (AS) PKCδ that is sensitive to specific kinase inhibitors and can be used to identify PKCδ substrates. AS PKCδ showed nearly 200 times higher affinity (Km) and 150 times higher efficiency (kcat/Km) than wild type (WT) PKCδ toward N6-(benzyl)-ATP. AS PKCδ was uniquely inhibited by 1-(tert-butyl)-3-(1-naphthyl)-1H-pyrazolo[3,4-d]pyrimidin-4-amine (1NA-PP1) and 1-(tert-butyl)-3-(2-methylbenzyl)-1H-pyrazolo[3,4-d]pyrimidin-4-amine (2MB-PP1) but not by other 4-amino-5-(4-methylphenyl)-7-(t-butyl)pyrazolo[3,4-d]pyrimidine (PP1) analogs tested, whereas WT PKCδ was insensitive to all PP1 analogs. To understand the mechanisms for specificity and affinity of these analogs, we created in silico WT and AS PKCδ homology models based on the crystal structure of PKCι. N6-(Benzyl)-ATP and ATP showed similar positioning within the purine binding pocket of AS PKCδ, whereas N6-(benzyl)-ATP was displaced from the pocket of WT PKCδ and was unable to interact with the glycine-rich loop that is required for phosphoryl transfer. The adenine rings of 1NA-PP1 and 2MB-PP1 matched the adenine ring of ATP when docked in AS PKCδ, and this interaction prevented the potential interaction of ATP with Lys-378, Glu-428, Leu-430, and Phe-633 residues. 1NA-PP1 failed to effectively dock within WT PKCδ. Other PP1 analogs failed to interact with either AS PKCδ or WT PKCδ. These results provide a structural basis for the ability of AS PKCδ to efficiently and specifically utilize N6-(benzyl)-ATP as a phosphate donor and for its selective inhibition by 1NA-PP1 and 2MB-PP1. Such homology modeling could prove useful in designing molecules to target PKCδ and other kinases to understand their function in cell signaling and to identify unique substrates.  相似文献   

14.
15.
Hypoxia promotes Na,K-ATPase endocytosis via protein kinase Cζ (PKCζ)-mediated phosphorylation of the Na,K-ATPase α subunit. Here, we report that hypoxia leads to the phosphorylation of 5′-AMP-activated protein kinase (AMPK) at Thr172 in rat alveolar epithelial cells. The overexpression of a dominant-negative AMPK α subunit (AMPK-DN) construct prevented the hypoxia-induced endocytosis of Na,K-ATPase. The overexpression of the reactive oxygen species (ROS) scavenger catalase prevented hypoxia-induced AMPK activation. Moreover, hypoxia failed to activate AMPK in mitochondrion-deficient ρ0-A549 cells, suggesting that mitochondrial ROS play an essential role in hypoxia-induced AMPK activation. Hypoxia-induced PKCζ translocation to the plasma membrane and phosphorylation at Thr410 were prevented by the pharmacological inhibition of AMPK or by the overexpression of the AMPK-DN construct. We found that AMPK α phosphorylates PKCζ on residue Thr410 within the PKCζ activation loop. Importantly, the activation of AMPK α was necessary for hypoxia-induced AMPK-PKCζ binding in alveolar epithelial cells. The overexpression of T410A mutant PKCζ prevented hypoxia-induced Na,K-ATPase endocytosis, confirming that PKCζ Thr410 phosphorylation is essential for this process. PKCζ activation by AMPK is isoform specific, as small interfering RNA targeting the α1 but not the α2 catalytic subunit prevented PKCζ activation. Accordingly, we provide the first evidence that hypoxia-generated mitochondrial ROS lead to the activation of the AMPK α1 isoform, which binds and directly phosphorylates PKCζ at Thr410, thereby promoting Na,K-ATPase endocytosis.When exposed to low oxygen levels (hypoxia), cells develop adaptative strategies to maintain adequate levels of ATP (21). These strategies include increasing the efficiency of energy-producing pathways, mostly through anaerobic glycolysis, while decreasing energy-consuming processes such as Na,K-ATPase activity (30). Alveolar hypoxia occurs in many respiratory disorders, and it has been shown to decrease epithelial active Na+ transport, leading to impaired fluid reabsorption (37, 41, 42). Active Na+ transport and, thus, alveolar fluid reabsortion are effected mostly via apical sodium channels and the basolateral Na,K-ATPase (32, 38, 42). We have reported previously that hypoxia inhibits Na,K-ATPase activity by promoting its endocytosis from the plasma membrane by a mechanism that requires the generation of mitochondrial reactive oxygen species (ROS) and the phosphorylation of the Na,K-ATPase α subunit at Ser18 by protein kinase Cζ (PKCζ) (8, 9).The 5′-AMP-activated protein kinase (AMPK) is a heterotrimeric Ser/Thr kinase composed of a catalytic α subunit and regulatory β and γ subunits. Both isoforms of the AMPK catalytic subunit (α1 and α2) form complexes with noncatalytic subunits. The α1 subunit is ubiquitously expressed, whereas the α2 subunit isoform is expressed predominantly in tissues like the liver, heart, and skeletal muscle (36). The α1 and α2 subunit isoforms have ∼90% homology in their N-terminal catalytic domains and ∼60% homology in their C-terminal domains (36), suggesting that they may have distinct downstream targets (31). AMPK activation requires phosphorylation at Thr172 in the activation loop of the α subunit by upstream kinases (12, 19). Findings from recent studies suggest that AMPK is an important signaling intermediary in coupling ion transport and metabolism (15). Indeed, it has been reported that the pharmacological activation of AMPK inhibits amiloride- and ouabain-sensitive epithelial Na+ transport (15). Moreover, the activities of the epithelial Na+ channel (ENaC) (2, 17), the Na,K-ATPase (40), and the cystic fibrosis transmembrane conductance regulator (17) have been shown to be inhibited by AMPK. Here, we provide evidence that hypoxia, via mitochondrial ROS, leads to AMPK activation and that AMPK binds to and directly phosphorylates PKCζ in an isoform-specific manner, thus promoting Na,K-ATPase endocytosis in alveolar epithelial cells (AEC).  相似文献   

16.
Radiation therapy for head and neck cancer can result in extensive damage to normal adjacent tissues such as the salivary gland and oral mucosa. We have shown previously that tyrosine phosphorylation at Tyr-64 and Tyr-155 activates PKCδ in response to apoptotic stimuli by facilitating its nuclear import. Here we have identified the tyrosine kinases that mediate activation of PKCδ in apoptotic cells and have explored the use of tyrosine kinase inhibitors for suppression of irradiation-induced apoptosis. We identify the damage-inducible kinase, c-Abl, as the PKCδ Tyr-155 kinase and c-Src as the Tyr-64 kinase. Depletion of c-Abl or c-Src with shRNA decreased irradiation- and etoposide-induced apoptosis, suggesting that inhibitors of these kinases may be useful therapeutically. Pretreatment with dasatinib, a broad spectrum tyrosine kinase inhibitor, blocked phosphorylation of PKCδ at both Tyr-64 and Tyr-155. Expression of “gate-keeper” mutants of c-Abl or c-Src that are active in the presence of dasatinib restored phosphorylation of PKCδ at Tyr-155 and Tyr-64, respectively. Imatinib, a c-Abl-selective inhibitor, also specifically blocked PKCδ Tyr-155 phosphorylation. Dasatinib and imatinib both blocked binding of PKCδ to importin-α and nuclear import, demonstrating that tyrosine kinase inhibitors can inhibit nuclear accumulation of PKCδ. Likewise, pretreatment with dasatinib also suppressed etoposide and radiation induced apoptosis in vitro. In vivo, pre-treatment of mice with dasatinib blocked radiation-induced apoptosis in the salivary gland by >60%. These data suggest that tyrosine kinase inhibitors may be useful prophylactically for protection of nontumor tissues in patients undergoing radiotherapy of the head and neck.  相似文献   

17.

Objective

Platelets play crucial roles in the pathophysiology of thrombosis and myocardial infarction. Protein kinase C ε (PKCε) is virtually absent in human platelets and its expression is precisely regulated during human megakaryocytic differentiation. On the basis of what is known on the role of platelet PKCε in other species, we hypothesized that platelets from myocardial infarction patients might ectopically express PKCε with a pathophysiological role in the disease.

Methods and Results

We therefore studied platelet PKCε expression from 24 patients with myocardial infarction, 24 patients with stable coronary artery disease and 24 healthy subjects. Indeed, platelets from myocardial infarction patients expressed PKCε with a significant frequency as compared to both stable coronary artery disease and healthy subjects. PKCε returned negative during patient follow-up. The forced expression of PKCε in normal donor platelets significantly increased their response to adenosine diphosphate-induced activation and adhesion to subendothelial collagen.

Conclusions

Our data suggest that platelet generations produced before the acute event retain PKCε-mRNA that is not down-regulated during terminal megakaryocyte differentiation. Results are discussed in the perspective of peri-infarctual megakaryocytopoiesis as a critical component of myocardial infarction pathophysiology.  相似文献   

18.
Neuroserpin is a member of the serine proteinase inhibitor superfamily. It can undergo a conformational transition to form polymers that are associated with the dementia familial encephalopathy with neuroserpin inclusion bodies and the wild-type protein can inhibit the toxicity of amyloid-β peptides in Alzheimer's disease. We have used a single molecule fluorescence method, two color coincidence detection, to determine the rate-limiting steps of the early stages of the polymerization of fluorophore-labeled neuroserpin and have assessed how this process is altered in the presence of Aβ1-40. Our data show that neuroserpin polymerization proceeds first by the unimolecular formation of an active monomer, followed by competing processes of both polymerization and formation of a latent monomer from the activated species. These data are not in keeping with the recently proposed domain swap model of polymer formation in which the latent species and activated monomer are likely to be formed by competing pathways directly from the unactivated monomeric serpin. Moreover, the Aβ1-40 peptide forms a weak complex with neuroserpin (dissociation constant of 10 ± 5 nM) that increases the amount of active monomer thereby increasing the rate of polymerization. The Aβ1-40 is displaced from the complex so that it acts as a catalyst and is not incorporated into neuroserpin polymers.  相似文献   

19.
20.
Protein kinase C (PKC) isoenzymes are multi-modular proteins activated at the membrane surface to regulate signal transduction processes. When activated by second messengers, PKC undergoes a drastic conformational and spatial transition from the inactive cytosolic state to the activated membrane-bound state. The complete structure of either state of PKC remains elusive. We demonstrate, using NMR spectroscopy, that the isolated Ca2+-sensing membrane-binding C2 domain of the conventional PKCα interacts with a conserved hydrophobic motif of the kinase C-terminal region, and we report a structural model of the complex. Our data suggest that the C-terminal region plays a dual role in regulating the PKC activity: activating, through sensitization of PKC to intracellular Ca2+ oscillations; and auto-inhibitory, through its interaction with a conserved positively charged region of the C2 domain.  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