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.
H Zhang  H Yu  X Wang  W Zheng  B Yang  J Pi  G He  W Qu 《PloS one》2012,7(8):e43004
α-Chlorohydrin is a common contaminant in food. Its (S)-isomer, (S)-α-chlorohydrin (SACH), is known for causing infertility in animals by inhibiting glycolysis of spermatozoa. The aim of present work was to examine the relationship between SACH and protein tyrosine phosphorylation (PTP), which plays a critical role in regulating mammalian sperm capacitation. In vitro exposure of SACH 50 μM to isolated rat epididymal sperm inhibited PTP. Sperm-specific glyceraldehyde 3-phosphate dehydrogenase (GAPDS) activities, the intracellular adenosine 5'-triphosphate (ATP) levels, 3'-5'-cyclic adenosine monophosphate (cAMP) levels and phosphorylation of protein kinase A (PKA) substrates in rat sperm were diminished dramatically, indicating that both glycolysis and the cAMP/PKA signaling pathway were impaired by SACH. The inhibition of both PTP and phosphorylation of PKA substrates by SACH could be restored by addition of cAMP analog dibutyryl-cAMP (dbcAMP) and phosphodiesterase inhibitor 3-isobutyl-1-methylxanthine (IBMX). Moreover, addition of glycerol protected glycolysis, ATP levels, phosphorylation of PKA substrates and PTP against the influence of SACH. These results suggested SACH inhibited PTP through blocking cAMP/PKA pathway in sperm, and PTP inhibition may play a role in infertility associated with SACH.  相似文献   

2.
Proper, graded communication between different cell types is essential for normal development and function. In the nervous system, heart, and for some cancer cells, part of this communication requires signaling by soluble and membrane-bound factors produced by the NRG1 gene. We have previously shown that glial-derived neurotrophic factors activate a rapid, localized release of soluble neuregulin from neuronal axons that can, in turn promote proper axoglial development (Esper, R. M., and Loeb, J. A. (2004) J. Neurosci. 24, 6218–6227). Here we elucidate the mechanism of this localized, regulated release by implicating the delta isoform of protein kinase C (PKC). Blocking the PKC delta isoform with either rottlerin, a selective antagonist, or small interference RNA blocks the regulated release of neuregulin from both transfected cells and primary neuronal cultures. PKC activation also leads to the rapid phosphorylation of the pro-NRG1 cytoplasmic tail on serine residues adjacent to the membrane-spanning segment, that, when mutated markedly reduce the rate of NRG1 activity release. These findings implicate this specific PKC isoform as an important factor for the cleavage and neurotrophin-regulated release of soluble NRG1 forms that have important effects in nervous system development and disease.The neuregulins (NRGs)2 are a family of growth and differentiation factors with a broad range of functions during development and in the adult. NRGs are necessary for glial and cardiac development and participate in a wide range of biologic processes ranging from proper formation of peripheral nerves and the neuromuscular junction to tumor growth (29). The NRGs have also been implicated as both potential mediators and therapeutic targets for a number of human diseases including cancer, schizophrenia, and multiple sclerosis (1012). NRGs function as mediators of cell-to-cell communication through a multitude of alternatively spliced isoforms arising from at least four distinct genes that bind to and activate members of the epidermal growth factor receptor family HER-2/3/4 (ErbB-2/3/4) (1319).Although all known isoforms of the NRG1 gene have an epidermal growth factor-like domain sufficient to bind to and activate its receptors (20), products of this gene are divided into three classes based on structurally and functionally different N-terminal regions (21) The type I and II forms have a unique N-terminal, heparin-binding Ig-like domain (2226). This Ig-like domain potentiates the biological activities of soluble NRG1 forms and leads to their highly selective tissue distributions through its affinity for specific cell-surface heparan sulfates (12, 20, 27, 28). These forms are first expressed as transmembrane precursors (pro-NRG1) that undergo proteolytic cleavage to release their soluble ectodomains. The type III NRG1 forms, on the other hand, are not typically released from cells, because their N-terminal domain consists of a cysteine-rich domain that can serve as a membrane tether making this form ideal for juxtacrine signaling. This form has been strongly implicated to be important peripheral nerve myelination (2931).While many of the biological functions of type I/II NRG1 forms are less clear, their ability to be released from axons in the peripheral and central nervous systems in a regulated manner provides the potential for long range cell-cell communication not possible from membrane-bound forms. Studies examining the regulation of type I NRG1 release from neuronal axons have implicated protein kinase C (PKC) as a mediator of NRG1 release from pro-NRG1 in transfected cell lines (32). Subsequent studies in intact neurons found that PKC activation was sufficient to release NRG1 from sensory and motor neuron axons and that NRG1 could also be released by Schwann cell-derived neurotrophic factors, such as BDNF and GDNF (1). Recently, the β-secretase protease BACE1 has been suggested to cleave these NRG1 forms so that when it is knocked out in mice, deficits similar to those seen in NRG1 knockouts are seen (33, 34). These findings suggest that reciprocal communication between NRG1s and neurotrophins could be an important mechanisms for local axoglial communication that is critical for normal peripheral nerve development. Consistently, PKC has been implicated as a key mediator for the electrically mediated release of NRG1 from cultured cerebellar granule cells and pontine nucleus neurons (35).The PKC family consists of 10 serine/threonine kinases isoforms (α, βI, βII, γ, δ, ϵ, ζ, θ, λ, and η) each with a unique cellular distribution, target specificity, mechanism of activation, and function (36). One of these functions promotes the cleavage and release of soluble signaling proteins that are initially synthesized as membrane-spanning precursors. In addition to NRG1, other proteins released upon PKC activation include epidermal growth factor, transforming growth factor-α, amyloid precursor protein, l-selectin, and interleukins (1, 3743). We hypothesize that neurotrophic factors induce the cleavage and release of NRG1 from pro-NRG1 through PKC activation. This hypothesis seems reasonable, because neurotrophin binding to the Trk family of neurotrophin receptor tyrosine kinases, but not the low affinity neurotrophin receptor p75 (44), activates phospholipase Cγ-mediated conversion of membrane-bound phosphatidylinositol bisphosphate to inositol triphosphate and diacylglycerol, which in turn, can activate PKC (4548). Although this can be achieved using phorbol 12-myristate 13-acetate (PMA), a diacylglycerol analog sufficient to activate most PKC isozymes (48), the exact PKC isoform and mechanism by which this occurs is not known. Here, we demonstrate NRG1 is released from cells through direct activation of the PKCδ isoform using siRNA and PKC isoform-specific inhibitors in transfected Chinese hamster ovary (CHO) cells, PC12, and primary neuronal cultures. We further demonstrate that PKC activation induces rapid phosphorylation of the cytoplasmic tail of pro-NRG1 on specific serine residues that are required for efficient NRG1 activity release. These findings provide mechanistic insights into how highly localized, reciprocal signaling occurs along neuronal axons, which has important implications for normal development and disease.  相似文献   

3.
4.
Vascular smooth muscle cell (VSMC) tone is regulated by the state of myosin light chain (MLC) phosphorylation, which is in turn regulated by the balance between MLC kinase and MLC phosphatase (MLCP) activities. RhoA activates Rho kinase, which phosphorylates the regulatory subunit of MLC phosphatase, thereby inhibiting MLC phosphatase activity and increasing contraction and vascular tone. Nitric oxide is an important mediator of VSMC relaxation and vasodilation, which acts by increasing cyclic GMP (cGMP) levels in VSMC, thereby activating cGMP-dependent protein kinase Iα (PKGIα). PKGI is known to phosphorylate Rho kinase, preventing Rho-mediated inhibition of MLC phosphatase, promoting vasorelaxation, although the molecular mechanisms that mediate this are unclear. Here we identify RhoA as a target of activated PKGIα and show further that PKGIα binds directly to RhoA, inhibiting its activation and translocation. In protein pulldown and immunoprecipitation experiments, binding of RhoA and PKGIα was demonstrated via a direct interaction between the amino terminus of RhoA (residues 1–44), containing the switch I domain of RhoA, and the amino terminus of PKGIα (residues 1–59), which includes a leucine zipper heptad repeat motif. Affinity assays using cGMP-immobilized agarose showed that only activated PKGIα binds RhoA, and a leucine zipper mutant PKGIα was unable to bind RhoA even if activated. Furthermore, a catalytically inactive mutant of PKGIα bound RhoA but did not prevent RhoA activation and translocation. Collectively, these results support that RhoA is a PKGIα target and that direct binding of activated PKGIα to RhoA is central to cGMP-mediated inhibition of the VSMC Rho kinase contractile pathway.  相似文献   

5.
The novel class of protein kinase C (nPKC) isoform η is expressed in platelets, but not much is known about its activation and function. In this study, we investigated the mechanism of activation and functional implications of nPKCη using pharmacological and gene knock-out approaches. nPKCη was phosphorylated (at Thr-512) in a time- and concentration-dependent manner by 2MeSADP. Pretreatment of platelets with MRS-2179, a P2Y1 receptor antagonist, or YM-254890, a Gq blocker, abolished 2MeSADP-induced phosphorylation of nPKCη. Similarly, ADP failed to activate nPKCη in platelets isolated from P2Y1 and Gq knock-out mice. However, pretreatment of platelets with P2Y12 receptor antagonist, AR-C69331MX did not interfere with ADP-induced nPKCη phosphorylation. In addition, when platelets were activated with 2MeSADP under stirring conditions, although nPKCη was phosphorylated within 30 s by ADP receptors, it was also dephosphorylated by activated integrin αIIbβ3 mediated outside-in signaling. Moreover, in the presence of SC-57101, a αIIbβ3 receptor antagonist, nPKCη dephosphorylation was inhibited. Furthermore, in murine platelets lacking PP1cγ, a catalytic subunit of serine/threonine phosphatase, αIIbβ3 failed to dephosphorylate nPKCη. Thus, we conclude that ADP activates nPKCη via P2Y1 receptor and is subsequently dephosphorylated by PP1γ phosphatase activated by αIIbβ3 integrin. In addition, pretreatment of platelets with η-RACK antagonistic peptides, a specific inhibitor of nPKCη, inhibited ADP-induced thromboxane generation. However, these peptides had no affect on ADP-induced aggregation when thromboxane generation was blocked. In summary, nPKCη positively regulates agonist-induced thromboxane generation with no effects on platelet aggregation.Platelets are the key cellular components in maintaining hemostasis (1). Vascular injury exposes subendothelial collagen that activates platelets to change shape, secrete contents of granules, generate thromboxane, and finally aggregate via activated αIIbβ3 integrin, to prevent further bleeding (2, 3). ADP is a physiological agonist of platelets secreted from dense granules and is involved in feedback activation of platelets and hemostatic plug stabilization (4). It activates two distinct G-protein-coupled receptors (GPCRs) on platelets, P2Y1 and P2Y12, which couple to Gq and Gi, respectively (58). Gq activates phospholipase Cβ (PLCβ), which leads to diacyl glycerol (DAG)2 generation and calcium mobilization (9, 10). On the other hand, Gi is involved in inhibition of cAMP levels and PI 3-kinase activation (4, 6). Synergistic activation of Gq and Gi proteins leads to the activation of the fibrinogen receptor integrin αIIbβ3. Fibrinogen bound to activated integrin αIIbβ3 further initiates feed back signaling (outside-in signaling) in platelets that contributes to the formation of a stable platelet plug (11).Protein kinase Cs (PKCs) are serine/threonine kinases known to regulate various platelet functional responses such as dense granule secretion and integrin αIIbβ3 activation (12, 13). Based on their structure and cofactor requirements, PKCs are divided in to three classes: classical (cofactors: DAG, Ca2+), novel (cofactors: DAG) and atypical (cofactors: PIP3) PKC isoforms (14). All the members of the novel class of PKC isoforms (nPKC), viz. nPKC isoforms δ, θ, η, and ε, are expressed in platelets (15), and they require DAG for activation. Among all the nPKCs, PKCδ (15, 16) and PKCθ (1719) are fairly studied in platelets. Whereas nPKCδ is reported to regulate protease-activated receptor (PAR)-mediated dense granule secretion (15, 20), nPKCθ is activated by outside-in signaling and contributes to platelet spreading on fibrinogen (18). On the other hand, the mechanism of activation and functional role of nPKCη is not addressed as yet.PKCs are cytoplasmic enzymes. The enzyme activity of PKCs is modulated via three mechanisms (14, 21): 1) cofactor binding: upon cell stimulus, cytoplasmic PKCs mobilize to membrane, bind cofactors such as DAG, Ca2+, or PIP3, release autoinhibition, and attain an active conformation exposing catalytic domain of the enzyme. 2) phosphorylations: 3-phosphoinositide-dependent kinase 1 (PDK1) on the membrane phosphorylates conserved threonine residues on activation loop of catalytic domain; this is followed by autophosphorylations of serine/threonine residues on turn motif and hydrophobic region. These series of phosphorylations maintain an active conformation of the enzyme. 3) RACK binding: PKCs in active conformation bind receptors for activated C kinases (RACKs) and are lead to various subcellular locations to access the substrates (22, 23). Although various leading laboratories have elucidated the activation of PKCs, the mechanism of down-regulation of PKCs is not completely understood.The premise of dynamic cell signaling, which involves protein phosphorylations by kinases and dephosphorylations by phosphatases has gained immense attention over recent years. PP1, PP2A, PP2B, PHLPP are a few of the serine/threonine phosphatases reported to date. Among them PP1 and PP2 phosphatases are known to regulate various platelet functional responses (24, 25). Furthermore, PP1c, is the catalytic unit of PP1 known to constitutively associate with αIIb and is activated upon integrin engagement with fibrinogen and subsequent outside-in signaling (26). Among various PP1 isoforms, recently PP1γ is shown to positively regulate platelet functional responses (27). Thus, in this study we investigated if the above-mentioned phosphatases are involved in down-regulation of nPKCη. Furthermore, reports from other cell systems suggest that nPKCη regulates ERK/JNK pathways (28). In platelets ERK is known to regulate agonist induced thromboxane generation (29, 30). Thus, we also investigated if nPKCη regulates ERK phosphorylation and thereby agonist-induced platelet functional responses.In this study, we evaluated the activation of nPKCη downstream of ADP receptors and its inactivation by an integrin-associated phosphatase PP1γ. We also studied if nPKCη regulates functional responses in platelets and found that this isoform regulates ADP-induced thromboxane generation, but not fibrinogen receptor activation in platelets.  相似文献   

6.
AMP-activated protein kinase (AMPK) is a heterotrimeric complex playing a crucial role in maintaining cellular energy homeostasis. Recently, homodimerization of mammalian AMPK and yeast ortholog SNF1 was shown by us and others. In SNF1, it involved specific hydrophobic residues in the kinase domain αG-helix. Mutation of the corresponding AMPK α-subunit residues (Val-219 and Phe-223) to glutamate reduced the tendency of the kinase to form higher order homo-oligomers, as was determined by the following three independent techniques in vitro: (i) small angle x-ray scattering, (ii) surface plasmon resonance spectroscopy, and (iii) two-dimensional blue native/SDS-PAGE. Recombinant protein as well as AMPK in cell lysates of primary cells revealed distinct complexes of various sizes. In particular, the assembly of very high molecular mass complexes was dependent on both the αG-helix-mediated hydrophobic interactions and kinase activation. In vitro and when overexpressed in double knock-out (α1−/−, α2−/−) mouse embryonic fibroblast cells, activation of mutant AMPK was impaired, indicating a critical role of the αG-helix residues for AMPK activation via its upstream kinases. Also inactivation by protein phosphatase 2Cα was affected in mutant AMPK. Importantly, activation of mutant AMPK by LKB1 was restored by exchanging the corresponding and conserved hydrophobic αG-helix residues of LKB1 (Ile-260 and Phe-264) to positively charged amino acids. These mutations functionally rescued LKB1-dependent activation of mutant AMPK in vitro and in cell culture. Our data suggest a physiological role for the hydrophobic αG-helix residues in homo-oligomerization of heterotrimers and cellular interactions, in particular with upstream kinases, indicating an additional level of AMPK regulation.The maintenance of energy homeostasis is a basic requirement of all living organisms. The AMP-activated protein kinase (AMPK)2 is crucially involved in this essential process by playing a central role in sensing and regulating energy metabolism on the cellular and whole body level (16). AMPK is also participating in several signaling pathways associated with cancer and metabolic diseases, like type 2 diabetes mellitus, obesity, and other metabolic disorders (79).Mammalian AMPK belongs to a highly conserved family of serine/threonine protein kinases with homologs found in all eukaryotic organisms examined (1, 3, 10). Its heterotrimeric structure includes a catalytic α-subunit and regulatory β- and γ-subunits. These subunits exist in different isoforms (α1, α2, β1, β2, γ1, γ2, and γ3) and splice variants (for γ2 and γ3) and can thus assemble to a broad variety of heterotrimeric isoform combinations. The α- and β-subunits possess multiple autophosphorylation sites, which have been implicated in regulation of subcellular localization and kinase activation (1115). The most critical step of AMPK activation, however, is phosphorylation of Thr-172 within the activation segment of the α-subunit kinase domain. At least two AMPK upstream kinases (AMPKKs) have been identified so far, namely the tumor suppressor kinase LKB1 in complex with MO25 and STRAD (16) and Ca2+/calmodulin-dependent protein kinase kinase-2 (CamKK2) (17). Furthermore, the transforming growth factor-β-activated kinase 1 was also shown to activate AMPK using a variety of in vitro approaches (18), but the physiological relevance of these findings remains unclear. Besides direct phosphorylation of Thr-172, AMPK activity is stimulated by the allosteric activator AMP, which can bind to two Bateman domains formed by two pairs of CBS domains within the γ-subunit (1922). Hereby bound AMP not only allosterically stimulates AMPK but also protects Thr-172 from dephosphorylation by protein phosphatase 2Cα (PP2Cα) and thus hinders inactivation of the kinase (19, 22, 23). Consequently, on the cellular level, AMPK is activated upon metabolic stress increasing the AMP/ATP ratio. Furthermore, AMPK activation can also be induced by several chemical compounds, like nucleoside 5-aminoimidazole-4-carboxamide-1-β-d-ribofuranoside (24) and the anti-diabetic drug Metformin (2528). In addition, the small molecule compound A-769662 was recently developed as a direct allosteric activator of AMPK (29, 30).Previous work in our groups proposed a model of AMPK regulation by AMP, which incorporates the major functional features and the latest structural information (31). The latter mainly included truncated core complexes of AMPK from different species (3235). Further valuable structural information is provided by the x-ray structures of the isolated catalytic domains, in particular of the human AMPK α2-subunit (Protein Data Bank code 2H6D) and its yeast ortholog SNF1 (36, 37). The kinase domain of SNF1 is capable of forming homodimers in the protein crystal, as well as in vitro in solution, in a unique way, which has not been observed previously in any other kinase (36). The dimer interface is predominantly formed by hydrophobic interactions of the loop-αG region, also known as subdomain X situated on the large kinase lobe (36, 38, 39), and it mainly involves Ile-257 and Phe-261. Because the T-loop activation segment was buried within the dimer interface, it was suggested that the dimeric state of the SNF1 catalytic domain represents the inactive form of the kinase. Intriguingly, it was shown in our groups by small angle x-ray scattering that AMPK self-organizes in a concentration-dependent manner to form homo-oligomers in solution (31). However, the interface responsible for oligomerization of the AMPK heterotrimer has remained elusive.Here we further investigate the distinct oligomeric states of the AMPK heterotrimer and suggest a possible regulatory function for this process. Most importantly, we provide conclusive evidence for participation of αG-helix residues in the recognition of AMPK by its upstream kinases LKB1 and CamKK2.  相似文献   

7.
Engagement of the T-cell receptor (TCR) in human primary T cells activates a cyclic AMP (cAMP)-protein kinase A (PKA)-Csk inhibitory pathway that prevents full T-cell activation in the absence of a coreceptor stimulus. Here, we demonstrate that stimulation of CD28 leads to recruitment to lipid rafts of a β-arrestin/phosphodiesterase 4 (PDE4) complex that serves to degrade cAMP locally. Redistribution of the complex from the cytosol depends on Lck and phosphatidylinositol 3-kinase (PI3K) activity. Protein kinase B (PKB) interacts directly with β-arrestin to form part of the supramolecular complex together with sequestered PDE4. Translocation is mediated by the PKB plextrin homology (PH) domain, thus revealing a new role for PKB as an adaptor coupling PI3K and cAMP signaling. Functionally, PI3K activation and phosphatidylinositol-(3,4,5)-triphosphate (PIP3) production, leading to recruitment of the supramolecular PKB/β-arrestin/PDE4 complex to the membrane via the PKB PH domain, results in degradation of the TCR-induced cAMP pool located in lipid rafts, thereby allowing full T-cell activation to proceed.T-cell receptor (TCR) stimulation alone is insufficient for activation of T cells, and sustainable T-cell immune responses require a second signal in addition to the TCR-mediated signal. The second signal is typically elicited by ligands B7-1 or B7-2 on antigen-presenting cells engaging the coreceptor CD28 to prevent anergy and apoptosis and enhancing interleukin-2 (IL-2) production and clonal expansion (4). Although CD28 plays a central role in T-cell activation in vivo (5), relatively little is known about the molecular basis for the increased efficacy of T-cell activation upon TCR and CD28 costimulation. Involvement of Lck, Itk, phosphatidylinositol 3-kinase (PI3K), SLP-76, Vav-1, and phospholipase C-γ (PLC-γ) has, however, been reported (43). CD28-mediated signals are transmitted via a short intracellular stretch in the receptor containing a conserved YMNM motif (32). Phosphorylation of Tyr173 in this motif by Lck and Fyn following CD28 ligation is key to efficient signal transduction (41), generating a binding site for the SH2 domain of the p85 regulatory subunit of PI3K (37, 40). CD28 may also contribute to TCR-dependent PI3K activity without recruiting PI3K directly (18). Whether engagement of CD28 alone can also induce PI3K activity has been a matter of controversy. However, recent reports confirming phosphorylation of the protein kinase B (PKB) substrate glycogen synthase kinase 3 (GSK3) upon CD28 ligation has demonstrated that this is indeed the case (6, 15). In addition, CD28 can recruit growth factor receptor-bound protein 2 (Grb2), and such association of Grb2 occurs via the phosphorylated YMNM motif as well as via the C-terminal PXXP motif (22, 35). The PXXP motif also binds and regulates Src family kinases (SFKs) (21, 47), and knock-in mice mutated in this motif were recently reported to have impaired IL-2 secretion (16).Ligation of the TCR induces cyclic AMP (cAMP) production (27). However, the significance of this observation is still not fully understood, as it is well established that cAMP potently inhibits T-cell function and proliferation (2, 45, 46, 50). The spatiotemporal dynamics of the activation-induced cAMP gradient also are not completely appreciated. We have previously shown that cAMP is rapidly produced in lipid rafts following engagement of the TCR in primary T cells (3). This activates a pool of PKA type I targeted to rafts by association with the anchoring protein Ezrin, forming part of a supramolecular complex where Ezrin, EBP50, and PAG provide a scaffold that is able to coordinate PKA phosphorylation and activation of Csk, thereby inhibiting T-cell activation (44, 50). In addition, we have demonstrated that CD3/CD28 costimulation leads to recruitment of type 4 phosphodiesterase (PDE4) isoforms to rafts, resulting in degradation of the TCR-induced cAMP pool (3). Thus, we envisage that TCR-induced cAMP production constitutes a negative feedback loop capable of abrogating T-cell activation in the absence of a second signal. In order then to allow full T-cell activation to proceed, cAMP-mediated inhibition must be lifted. This appears to occur in the presence of a costimulus involving CD28 acting to trigger recruitment of PDE4 to lipid rafts, thereby degrading cAMP at this spatially critical location and resulting in an overriding positive feed-forward signal rather than the negative feedback loop activated from the TCR. In addition, a recent publication by Conche et al. has also found a possible stimulatory effect of cAMP, as the paper surprisingly showed that a transient cAMP increase shortly after TCR triggering may potentiate the calcium component of the TCR signaling. This could constitute a positive feed-forward in addition to the negative feedback signal by cAMP (12).Spatial organization and recruitment of mediators of specific pathways as outlined above are essential to ensure signaling specificity and amplification. Among the many protein scaffolds linking effector molecules into linear pathways, β-arrestins have been reported to confer cross talk with a growing list of molecules important in cellular trafficking and signal transduction, including Src family members and mitogen-activated protein (MAP) kinases (reviewed in reference 14). The arrestins were first identified as having a role in desensitization of G protein-coupled receptors (GPCRs) (9); later, they were discovered to be involved in receptor internalization by interacting with clathrin and AP-2, thereby bringing activated receptors to clathrin-coated pits for endocytosis (19, 26). A role for β-arrestin in the spatially localized degradation of cAMP by scaffolding PDE4 isoforms to the proximity of cAMP generation at the plasma membrane has also been suggested (3, 7, 30, 38).In the present study, we uncover a novel pathway that defines how T-cell costimulation elicits recruitment of PDE4 to lipid rafts to overcome cAMP-mediated inhibition of T-cell activation. This pathway is initiated by CD28 engagement leading to PI3K activation and phosphatidylinositol-(3,4,5)-triphosphate (PIP3) production and resulting in recruitment of a supramolecular complex of PKB/β-arrestin/PDE4 targeted to the plasma membrane due to sequestration via the PKB plextrin homology (PH) domain. Functionally, this pathway is essential for CD28 costimulation to strengthen and sustain T-cell immune responses.  相似文献   

8.
We took a discovery approach to explore the actions of cAMP and two of its analogs, one a cAMP mimic ((Sp)-adenosine cyclic 3′:5′-monophosphorothioate ((Sp)-cAMPS)) and the other a diastereoisomeric antagonist ((Rp)-cAMPS), on a model system of the type Iα cyclic AMP-dependent protein kinase holoenzyme, RIα(91–244)·C-subunit, by using fluorescence spectroscopy and amide H/2H exchange mass spectrometry. Specifically, for the fluorescence experiments, fluorescein maleimide was conjugated to three cysteine single residue substitution mutants, R92C, T104C, and R239C, of RIα(91–244), and the effects of cAMP, (Sp)-cAMPS, and (Rp)-cAMPS on the kinetics of R-C binding and the time-resolved anisotropy of the reporter group at each conjugation site were measured. For the amide exchange experiments, ESI-TOF mass spectrometry with pepsin proteolytic fragmentation was used to assess the effects of (Rp)-cAMPS on amide exchange of the RIα(91–244)·C-subunit complex. We found that cAMP and its mimic perturbed at least parts of the C-subunit interaction Sites 2 and 3 but probably not Site 1 via reduced interactions of the linker region and αC of RIα(91–244). Surprisingly, (Rp)-cAMPS not only increased the affinity of RIα(91–244) toward the C-subunit by 5-fold but also produced long range effects that propagated through both the C- and R-subunits to produce limited unfolding and/or enhanced conformational flexibility. This combination of effects is consistent with (Rp)-cAMPS acting by enhancing the internal entropy of the R·C complex. Finally, the (Rp)-cAMPS-induced increase in affinity of RIα(91–244) toward the C-subunit indicates that (Rp)-cAMPS is better described as an inverse agonist because it decreases the fractional dissociation of the cyclic AMP-dependent protein kinase holoenzyme and in turn its basal activity.Cyclic AMP-dependent protein kinase (PKA)1 plays a crucial role in a plethora of cellular functions. All isoforms of PKA are composed of two catalytic (C) subunits and homodimeric regulatory (R) subunits (13). As the name implies, cAMP is a major PKA regulator (4). Much progress has been made in the last decade in delineating the molecular basis of action of cAMP. An important tactic in this endeavor has been through the comparison of the effects of cAMP with those of two phosphorothioate cAMP analogs: (Sp)-cAMPS (a cAMP mimic) and (Rp)-cAMPS (an antagonist and a diastereoisomer of (Sp)-cAMPS). Although the importance of geometry of the sulfur substitution is critical in determining the pharmacological properties of the two phosphorothioate cAMP analogs, the molecular basis for this behavior is not fully understood. To date, these comparisons have only been made using either wild-type or truncated mutants of the type Iα regulatory subunit (RIα) that are free in solution, not complexed to the C-subunit. X-ray spectroscopic examination of ligand-bound RIα(92–379) complexes reveals few differences between ligand-bound complexes, but the (Rp)-cAMPS complex is structurally “looser” with higher thermal factors than complexes formed with either cAMP or (Sp)-cAMPS (5). This is consistent with the observation that both cAMP and (Sp)-cAMPS, but not (Rp)-cAMPS, raise the urea concentration required for wild-type RIα unfolding (6). Further insight into the structural basis of cAMP action stems from NMR spectroscopic comparison of the effects of (Rp)-cAMPS, cAMP, and (Sp)-cAMPS on chemical shifts and 15N relaxation of the RIα(119–244) mutant (7). In addition to producing fewer significant chemical shift changes than either cAMP or (Sp)-cAMPS, (Rp)-cAMPS binding is associated with enhanced millisecond to microsecond time scale backbone motions of a β-turn (β2,3 loop) and around the phosphate-binding cassette (PBC) (7).Further insight into the molecular basis of actions of cAMP and its analogs should come from the analysis of ligand-bound R·C complexes. Unfortunately, the large size of even the heterodimeric R·C complex (∼95 kDa) and the difficulty of preparing (Rp)-cAMPS·R·C-subunit crystals currently preclude the use of both NMR spectroscopy and x-ray crystallography. Consequently, we took two alternative lower resolution approaches to this issue. One approach involves the use of site-directed labeling combined with fluorescence spectroscopy to examine both the effects of cAMP and its analogs on R-C subunit binding kinetics and on the conformational dynamics of RIα(91–244). RIα(91–244) includes the “A” cyclic nucleotide binding (CNB) domain, the pseudosubstrate, and linker domains and represents the minimal segments necessary for high affinity C-subunit binding (Fig. 1) (8). The other approach involves an examination of the effects of cAMP and its analogs on solvent exposure/conformational flexibility of RIα(91–244)·C-subunit complex using H/2H amide exchange measured with a combination of mass spectrometry (ESI-Q-TOF) and proteolytic fragmentation. In the first approach, fluorescein maleimide (FM) was conjugated to three cysteine substitution mutants with the substitution sites located near or within the pseudosubstrate sequence, the linker domain, or αC (R92C, T104C, and R239C, respectively) of RIα(91–244) (Fig. 1). The time-resolved fluorescence anisotropy results suggest that cAMP and (Sp)-cAMPS reduce the interaction of the RIα linker domain and αC with the two peripheral R-C interaction sites on the C-subunit (so-called Sites 2 and 3) without affecting the interaction of the pseudosubstrate sequence with the active site cleft (so-called Site 1). Because of limitations of the amide H/2H exchange experiments, only the effects of (Rp)-cAMPS on H/2H amide exchange in RIα(91–244)·C-subunit complex could be investigated. The results showed that (Rp)-cAMPS induces a relatively widespread increase in amide exchange, indicating limited unfolding and/or enhanced conformational flexibility that is propagated almost globally through the C-subunit and, at least, part of RIα. These conformational changes were accompanied by a 5-fold increase in the affinity of RIα(91–244) toward C-subunit, suggesting that, at least, some of the (Rp)-cAMPS effects are mediated by an increase in internal entropy. Finally, the (Rp)-cAMPS-induced increase in R-C affinity indicates that (Rp)-cAMPS is better described as an inverse agonist because the basal activity of the PKA holoenzyme should be decreased by (Rp)-cAMPS.Open in a separate windowFig. 1.Overview of PKA structure and cAMP analogs. A, domain organization of RIα showing the domain boundaries of RIα(91–244) where the pseudosubstrate in green is connected to CNB-A domain in blue by a linker segment. B, structure of RIα(91–244) in the C-subunit-bound conformation (Protein Data Bank code 1U7E (23)) showing the pseudosubstrate in green, linker in yellow, and helical subdomain comprising helices αN, αA, αB, and αC in blue and β-subdomain in tan. The PBC is in red. C, structure of the C·RIα(91–244) holoenzyme showing the C-subunit in tan and RIα(91–244) in blue. Sites for introduction of cysteines by site-directed mutagenesis are represented by red circles. The cAMP binding site (PBC) is in red. D, structure of cAMP showing the 2′-OH group and 3′–5′ phosphodiester bond. The exocyclic oxygens upon replacement with sulfur atoms to generate the (Sp)-cAMPS and (Rp)-cAMPS diastereomers are highlighted.  相似文献   

9.
Atypical protein kinase C (PKC) ζ is an important regulator of inflammation through activation of the nuclear factor-κB (NF-κB) pathway. Chromatin remodeling on pro-inflammatory genes plays a pivotal role in cigarette smoke (CS)- and lipopolysaccharide (LPS)-induced abnormal lung inflammation. However, the signaling mechanism whereby chromatin remodeling occurs in CS- and LPS-induced lung inflammation is not known. We hypothesized that PKCζ is an important regulator of chromatin remodeling, and down-regulation of PKCζ ameliorates lung inflammation by CS and LPS exposures. We determined the role and molecular mechanism of PKCζ in abnormal lung inflammatory response to CS and LPS exposures in PKCζ-deficient (PKCζ−/−) and wild-type mice. Lung inflammatory response was decreased in PKCζ−/− mice compared with WT mice exposed to CS and LPS. Moreover, inhibition of PKCζ by a specific pharmacological PKCζ inhibitor attenuated CS extract-, reactive aldehydes (present in CS)-, and LPS-mediated pro-inflammatory mediator release from macrophages. The mechanism underlying these findings is associated with decreased RelA/p65 phosphorylation (Ser311) and translocation of the RelA/p65 subunit of NF-κB into the nucleus. Furthermore, CS/reactive aldehydes and LPS exposures led to activation and translocation of PKCζ into the nucleus where it forms a complex with CREB-binding protein (CBP) and acetylated RelA/p65 causing histone phosphorylation and acetylation on promoters of pro-inflammatory genes. Taken together, these data suggest that PKCζ plays an important role in CS/aldehyde- and LPS-induced lung inflammation through acetylation of RelA/p65 and histone modifications via CBP. These data provide new insights into the molecular mechanisms underlying the pathogenesis of chronic inflammatory lung diseases.  相似文献   

10.
11.
The protein kinase PKR is activated by RNA with double-stranded (ds) structure and subsequently impairs translation through phosphorylation of protein synthesis initiation factor eIF-2α. PKR also mediates activation of signal transduction pathways leading to interferon beta (IFN-β) gene induction following virus-infection or RNA transfection. We previously demonstrated in measles virus-infected cells that PKR is required for the maximal induction of IFN-β gene expression by the interferon promoter stimulator gene 1 (IPS-1) adaptor-dependent cytosolic RNA sensor pathway. While both IPS-1 and PKR are important mediators of IFN-β induction, with PKR contributing to an enhanced NF-κB activation, the mechanism by which PKR enhances NF-κB activity and amplifies IFN-β induction is unresolved. Herein we tested the possibility that PKR could activate signal transduction pathways indirectly through translational control responses. Following transfection with synthetic or natural dsRNAs or infection with measles virus, we observed increased mRNA but decreased protein levels for the inhibitor of NF-κB signaling, IκB-α, that correlated with PKR activation and eIF-2α phosphorylation. Importantly, knockdown of PKR increased IκB-α protein levels and impaired IFN-β induction. Additionally, inhibition of translation by cycloheximide treatment rescued IFN-β induction following PKR knockdown but not IPS-1 knockdown. Mutation of eIF-2α to prevent phosphorylation also impaired IFN-β induction in PKR-sufficient virus-infected cells. These results suggest that an eIF-2α-dependent translation inhibition mechanism is sufficient to explain the PKR-mediated amplification of IPS-1-dependent IFN-β induction by foreign RNA.  相似文献   

12.
Pancreatic β-cells secrete insulin in response to metabolic and hormonal signals to maintain glucose homeostasis. Insulin secretion is under the control of ATP-sensitive potassium (KATP) channels that play key roles in setting β-cell membrane potential. Leptin, a hormone secreted by adipocytes, inhibits insulin secretion by increasing KATP channel conductance in β-cells. We investigated the mechanism by which leptin increases KATP channel conductance. We show that leptin causes a transient increase in surface expression of KATP channels without affecting channel gating properties. This increase results primarily from increased channel trafficking to the plasma membrane rather than reduced endocytosis of surface channels. The effect of leptin on KATP channels is dependent on the protein kinases AMP-activated protein kinase (AMPK) and PKA. Activation of AMPK or PKA mimics and inhibition of AMPK or PKA abrogates the effect of leptin. Leptin activates AMPK directly by increasing AMPK phosphorylation at threonine 172. Activation of PKA leads to increased channel surface expression even in the presence of AMPK inhibitors, suggesting AMPK lies upstream of PKA in the leptin signaling pathway. Leptin signaling also leads to F-actin depolymerization. Stabilization of F-actin pharmacologically occludes, whereas destabilization of F-actin simulates, the effect of leptin on KATP channel trafficking, indicating that leptin-induced actin reorganization underlies enhanced channel trafficking to the plasma membrane. Our study uncovers the signaling and cellular mechanism by which leptin regulates KATP channel trafficking to modulate β-cell function and insulin secretion.  相似文献   

13.
IL-33 contributes to disease processes in association with Th1 and Th2 phenotypes. IL-33 mRNA is rapidly regulated, but the fate of synthesized IL-33 protein is unknown. To understand the interplay among IL-33, IFN-γ, and IL-4 proteins, recombinant replication-deficient adenoviruses were produced and used for dual expression of IL-33 and IFN-γ or IL-33 and IL-4. The effects of such dual gene delivery were compared with the effects of similar expression of each of these cytokines alone. In lung fibroblast culture, co-expression of IL-33 and IFN-γ resulted in suppression of the levels of both proteins, whereas co-expression of IL-33 and IL-4 led to mutual elevation. In vivo, co-expression of IL-33 and IFN-γ in the lungs led to attenuation of IL-33 protein levels. Purified IFN-γ also attenuated IL-33 protein in fibroblast culture, suggesting that IFN-γ controls IL-33 protein degradation. Specific inhibition of caspase-1, -3, and -8 had minimal effect on IFN-γ-driven IL-33 protein down-regulation. Pharmacological inhibition, siRNA-mediated silencing, or gene deficiency of STAT1 potently up-regulated IL-33 protein expression levels and attenuated the down-regulating effect of IFN-γ on IL-33. Stimulation with IFN-γ strongly elevated the levels of the LMP2 proteasome subunit, known for its role in IFN-γ-regulated antigen processing. siRNA-mediated silencing of LMP2 expression abrogated the effect of IFN-γ on IL-33. Thus, IFN-γ, IL-4, and IL-33 are engaged in a complex interplay. The down-regulation of IL-33 protein levels by IFN-γ in pulmonary fibroblasts and in the lungs in vivo occurs through STAT1 and non-canonical use of the LMP2 proteasome subunit in a caspase-independent fashion.  相似文献   

14.
A plant's capability to cope with environmental challenges largely relies on signal transmission through mitogen-activated protein kinase (MAPK) cascades. In Arabidopsis thaliana, MPK3 is particularly strongly associated with numerous abiotic and biotic stress responses. Identification of MPK3 substrates is a milestone towards improving stress resistance in plants. Here, we characterize AZI1, a lipid transfer protein (LTP)-related hybrid proline-rich protein (HyPRP), as a novel target of MPK3. AZI1 is phosphorylated by MPK3 in vitro. As documented by co-immunoprecipitation and bimolecular fluorescence complementation experiments, AZI1 interacts with MPK3 to form protein complexes in planta. Furthermore, null mutants of azil are hypersensitive to salt stress, while AZIl-overexpressing lines are markedly more tolerant. AZI1 overexpression in the mpk3 genetic background partially alleviates the salt-hypersensitive phenotype of this mutant, but functional MPK3 appears to be required for the full extent of AZIl-conferred robustness. Notably, this robustness does not come at the expense of normal development. Immunoblot and RT-PCR data point to a role of MPK3 as positive regulator of AZI1 abundance.  相似文献   

15.
Many agents that activate hematopoietic cells use phos pha tidyl ino si tol 3,4,5-trisphosphate (PtdIns 3,4,5-P3) to initiate signaling cascades. The SH2 domain-containing inositol 5′ phosphatase, SHIP1, regulates hematopoietic cell function by opposing the action of phos pha tidyl ino si tol 3-kinase and reducing the levels of PtdIns 3,4,5-P3. Activation of the cyclic AMP-de pend ent protein kinase (PKA) also opposes many of the pro-inflammatory responses of hematopoietic cells. We tested to see whether the activity of SHIP1 was regulated via phos pho ryl a tion with PKA. We prepared pure recombinant SHIP1 from HEK-293 cells and found it can be rapidly phos pho ryl a ted by PKA to a stoichiometry of 0.6 mol of PO4/mol of SHIP1. In 32P-labeled HEK-293 cells transfected with SHIP1, stimulation with Sp-adenosine 3′,5′-cyclic monophosphorothioate triethylammonium salt hydrate (Sp-cAMPS) or activation of the β-adrenergic receptor increased the phos pho ryl a tion state of SHIP1. Inhibition of protein phosphatase activity with okadaic acid also increased the phos pho ryl a tion of SHIP1. Phosphorylation of SHIP1 in vitro or in cells by PKA increased the 5′ phosphatase activity of SHIP1 by 2–3-fold. Elevation of Ca2+ in DT40 cells in response to B cell receptor cross-linking, an indicator of PtdIns 3,4,5-P3 levels, was markedly blunted by pretreatment with Sp-cAMPS. This effect was absent in SHIP−/− DT40 cells showing that the effect of Sp-cAMPS in DT40 cells is SHIP1-de pend ent. Sp-cAMPS also blunted the ability of the B cell receptor to increase the phos pho ryl a tion of Akt in DT40 and A20 cells. Overall, activation of G protein-coupled receptors that raise cyclic AMP cause SHIP1 to be phos pho ryl a ted and stimulate its inositol phosphatase activity. These results outline a novel mechanism of SHIP1 regulation.Activation of phosphatidylinositol 3-kinase (PtdIns 3-kinase)2 is central to regulation of multiple cell functions including cell shape changes, cell migration, cell activation, and proliferation (1). PtdIns 3-kinase phosphorylates phosphatidylinositol 4,5-bisphosphate in the inner leaflet of the plasma membrane to generate phosphatidylinositol 3,4,5-trisphosphate (PtdIns 3,4,5-P3) (2). PtdIns 3,4,5-P3 then activates downstream signaling pathways by interacting with pleckstrin homology domain-containing proteins, such as phosphoinositide-dependent kinase 1 and the serine-threonine kinase Akt (3). The finding of abnormal activation of the PtdIns 3-kinase pathway in cancer cells has led to interest in the development of inhibitors for PtdIns 3-kinase (4).The level of PtdIns 3,4,5-P3 is stimulated by multiple members of the PtdIns 3-kinase family (2) and is opposed by two phosphatidylinositol phosphatases: the Src homology 2 (SH2) domain-containing inositol 5′ phosphatase (SHIP) and the 3′ inositol phosphatase, phosphatase and tensin homolog (PTEN) (5). PTEN removes phosphate from the 3′ position in the inositol ring of PtdIns 3,4,5-P3 and converts it to phosphatidylinositol 4,5-bisphosphate (6). PTEN has a C2 domain, a PDZ-binding motif, and a N-terminal phosphatidylinositol 4,5-bisphosphate binding motif essential for translocation to the membrane and interaction with other regulatory proteins (7). There are serine and threonine residues in PTEN that have been found to be phosphorylated, but their role in regulating the activity of the enzyme is not clear (8). Mutations in the PTEN protein have been observed in many tumors, suggesting a role for this enzyme in cancer (9).In contrast, SHIP dephosphorylates the 5′ position on the inositol ring and produces phosphatidylinositol 3,4-bisphosphate (10). There are three isoforms of SHIP: the 145-kDa hematopoietic cell restricted SHIP (also known as SHIP1); the 104-kDa stem cell-restricted SHIP, sSHIP; and the more widely expressed 150-kDa SHIP2 (11). SHIP1 is the major inositol phosphatase regulating PtdIns 3,4,5-P3 in monocytes, macrophages, B cells, and T cells (11). SHIP1 has three known structural features: the N-terminal SH2 domain, the central inositol 5′ phosphatase domain, and two NPXY sequences in the C-terminal region. The currently accepted model for regulation of PtdIns 3,4,5-P3 levels by SHIP1 envisions translocation of SHIP1 from the cytosol to the membrane. Upon stimulation by growth factors, cytokine receptors, or immunoreceptors, SHIP1 is recruited via its N-terminal SH2 domain to phosphorylated tyrosine residues in receptor kinases and degrades the elevated levels of PtdIns 3,4,5-P3 near the activated receptor (12). During this translocation process, SHIP1 is not thought to change its 5′ phosphatase activity (13). Although it is known that SHIP1 can be phosphorylated on tyrosine residues by the lyn cytoplasmic kinase (12) or following the activation of the T cell receptor (14), neither event appears to influence the 5′ phosphatase activity. To date, direct regulation of SHIP1 activity by serine/threonine kinases has not been studied.Activation of G protein-coupled receptors that raise cAMP (i.e. β-adrenergic receptors or adenosine A2a receptors) is known to blunt the pro-inflammatory responses generated by receptors that raise the level of PtdIns 3,4,5-P3 (15). Therefore, we investigated the possibility that phosphorylation of SHIP1 by cyclic AMP-dependent protein kinase (PKA) might regulate the activity of SHIP1. We found that SHIP1 can be phosphorylated by PKA both in vitro and in cells leading to a stimulation of SHIP1 activity. Activation of PKA in DT40 and A20 cells blunted indicators of the PtdIns 3,4,5-P3 response to B cell receptor stimulation. These results indicate that SHIP1 activity can be regulated both in vitro and in cells by activation of the cyclic AMP-dependent protein kinase and highlight a new mode of SHIP regulation by G protein-coupled receptors.  相似文献   

16.
17.
Discrimination between self and non-self is a prerequisite for any defence mechanism; in innate defence, this discrimination is often mediated by lectins recognizing non-self carbohydrate structures and so relies on an arsenal of host lectins with different specificities towards target organism carbohydrate structures. Recently, cytoplasmic lectins isolated from fungal fruiting bodies have been shown to play a role in the defence of multicellular fungi against predators and parasites. Here, we present a novel fruiting body lectin, CCL2, from the ink cap mushroom Coprinopsis cinerea. We demonstrate the toxicity of the lectin towards Caenorhabditis elegans and Drosophila melanogaster and present its NMR solution structure in complex with the trisaccharide, GlcNAcβ1,4[Fucα1,3]GlcNAc, to which it binds with high specificity and affinity in vitro. The structure reveals that the monomeric CCL2 adopts a β-trefoil fold and recognizes the trisaccharide by a single, topologically novel carbohydrate-binding site. Site-directed mutagenesis of CCL2 and identification of C. elegans mutants resistant to this lectin show that its nematotoxicity is mediated by binding to α1,3-fucosylated N-glycan core structures of nematode glycoproteins; feeding with fluorescently labeled CCL2 demonstrates that these target glycoproteins localize to the C. elegans intestine. Since the identified glycoepitope is characteristic for invertebrates but absent from fungi, our data show that the defence function of fruiting body lectins is based on the specific recognition of non-self carbohydrate structures. The trisaccharide specifically recognized by CCL2 is a key carbohydrate determinant of pollen and insect venom allergens implying this particular glycoepitope is targeted by both fungal defence and mammalian immune systems. In summary, our results demonstrate how the plasticity of a common protein fold can contribute to the recognition and control of antagonists by an innate defence mechanism, whereby the monovalency of the lectin for its ligand implies a novel mechanism of lectin-mediated toxicity.  相似文献   

18.
Molecular mechanisms of plasticity at GABAergic synapses are currently poorly understood. To identify signaling cascades that converge onto GABAergic postsynaptic density proteins, we performed MS analysis using gephyrin isolated from rat brain and identified multiple novel phosphorylation and acetylation residues on gephyrin. Here, we report the characterization of one of these phosphoresidues, Ser-268, which when dephosphorylated leads to the formation of larger postsynaptic scaffolds. Using a combination of mutagenesis, pharmacological treatment, and biochemical assays, we identify ERK as the kinase phosphorylating Ser-268 and describe a functional interaction between residues Ser-268 and Ser-270. We further demonstrate that alterations in gephyrin clustering via ERK modulation are reflected by amplitude and frequency changes in miniature GABAergic postsynaptic currents. We unravel novel mechanisms for activity- and ERK-dependent calpain action on gephyrin, which are likely relevant in the context of cellular signaling affecting GABAergic transmission and homeostatic synaptic plasticity in pathology.  相似文献   

19.
20.
Cyclic AMP (cAMP) and cyclic GMP (cGMP) suppress apoptosis in many cell types, including hepatocytes. We have previously shown that membrane-permeable cAMP and cGMP analogs attenuate tumor necrosis factor α plus actinomycin D (TNFα/ActD)-induced apoptosis in hepatocytes at a step upstream of caspase activation and cytochrome c release. Recently we have also shown that FADD levels increase 10 folds in response to TNFα/ActD. Therefore we hypothesized that cAMP and cGMP would inhibit FADD upregulation. We show here that cyclic nucleotide analogs dibutyryl cAMP (db-cAMP) and 8-bromo-cGMP (Br-cGMP) inhibit cell death and the cleavages of multiple caspases including caspase-10, -9, -8, -3, and -2, as well as suppress FADD protein up-regulation in TNFα/ActD-induced apoptosis. The inhibitory effects of cAMP were seen at lower concentrations than cGMP. Both cAMP and cGMP prevented FADD overexpression and cell death in hepatocytes transfected with the FADD gene. A protein kinase A (PKA) inhibitor, KT 5720, reversed the inhibition of FADD protein levels induced by cAMP or cGMP. In conclusion, our findings indicate that cAMP and cGMP prevent TNFα/ActD-induced apoptosis in hepatocytes and that this occurs in association with a near complete inhibition of the upregulation of FADD via a PKA-dependent mechanism. Supported by the National Institutes of Health Grant GM-44100 (to T.R.B).  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