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 46 毫秒
1.
Calcium (Ca2+) signaling by the pro-inflammatory cytokine interleukin-1 (IL-1) is dependent on focal adhesions, which contain diverse structural and signaling proteins including protein phosphatases. We examined here the role of protein-tyrosine phosphatase (PTP) α in regulating IL-1-induced Ca2+ signaling in fibroblasts. IL-1 promoted recruitment of PTPα to focal adhesions and endoplasmic reticulum (ER) fractions, as well as tyrosine phosphorylation of the ER Ca2+ release channel IP3R. In response to IL-1, catalytically active PTPα was required for Ca2+ release from the ER, Src-dependent phosphorylation of IP3R1 and accumulation of IP3R1 in focal adhesions. In pulldown assays and immunoprecipitations PTPα was required for the association of PTPα with IP3R1 and c-Src, and this association was increased by IL-1. Collectively, these data indicate that PTPα acts as an adaptor to mediate functional links between focal adhesions and the ER that enable IL-1-induced Ca2+ signaling.The interleukin-1 (IL-1)3 family of pro-inflammatory cytokines mediates host responses to infection and injury. Impaired control of IL-1 signaling leads to chronic inflammation and destruction of extracellular matrices (1, 2), as seen in pathological conditions such as pulmonary fibrosis (3), rheumatoid arthritis (4, 5), and periodontitis (6). IL-1 elicits multiple signaling programs, some of which trigger Ca2+ release from the endoplasmic reticulum (ER) as well as expression of multiple cytokines and inflammatory factors including c-Fos and c-Jun (7, 8), and matrix metalloproteinases (9, 10), which mediate extracellular matrix degradation via mitogen-activated protein kinase-regulated pathways (11).In anchorage-dependent cells including fibroblasts and chondrocytes, focal adhesions (FAs) are required for IL-1-induced Ca2+ release from the ER and activation of ERK (1214). FAs are actin-enriched adhesive domains composed of numerous (>50) scaffolding and signaling proteins (1517). Many FA proteins are tyrosine-phosphorylated, including paxillin, focal adhesion kinase, and src family kinases, all of which are crucial for the assembly and disassembly of FAs (1821). Protein-tyrosine phosphorylation plays a central role in regulating many cellular processes including adhesion (22, 23), motility (24), survival (25), and signal transduction (2629). Phosphorylation of proteins by kinases is balanced by protein-tyrosine phosphatases (PTP), which can enhance or attenuate downstream signaling by dephosphorylation of tyrosine residues (3032).PTPs can be divided into two main categories: receptor-like and intracellular PTPs (33). Two receptor-like PTPs have been localized to FA (leukocyte common antigen-related molecule and PTPα). Leukocyte common antigen-related molecule can dephosphorylate and mediate degradation of p130cas, which ultimately leads to cell death (34, 35). PTPα contains a heavily glycosylated extracellular domain, a transmembrane domain, and two intracellular phosphatase domains (33, 36). The amino-terminal domain predominantly mediates catalytic activity, whereas the carboxyl-terminal domain serves a regulatory function (37, 38). PTPα is enriched in FA (23) and is instrumental in regulating FA dynamics (39) via activation of c-Src/Fyn kinases by dephosphorylating the inhibitory carboxyl tyrosine residue, namely Tyr529 (22, 4042) and facilitation of integrin-dependent assembly of Src-FAK and Fyn-FAK complexes that regulate cell motility (43). Although PTPα has been implicated in formation and remodeling of FAs (44, 45), the role of PTPα in FA-dependent signaling is not defined.Ca2+ release from the ER is a critical step in integrin-dependent IL-1 signal transduction and is required for downstream activation of ERK (13, 46). The release of Ca2+ from the ER depends on the inositol 1,4,5-triphosphate receptor (IP3R), which is an IP3-gated Ca2+ channel (47). All of the IP3R subtypes (subtypes 1–3) have been localized to the ER, as well as other the plasma membrane and other endomembranes (4850). Further, IP3R may associate with FAs, enabling the anchorage of the ER to FAs (51, 52). However, the molecule(s) that provide the structural link for this association has not been defined.FA-restricted, IL-1-triggered signal transduction in anchorage-dependent cells may rely on interacting proteins that are enriched in FAs and the ER (53). Here, we examined the possibility that PTPα associates with c-Src and IP3R to functionally link FAs to the ER, thereby enabling IL-1 signal transduction.  相似文献   

2.
3.
Fertilization induces a species-specific Ca2+ transient with specialized spatial and temporal dynamics, which are essential to temporally encode egg activation events such as the block to polyspermy and resumption of meiosis. Eggs acquire the competence to produce the fertilization-specific Ca2+ transient during oocyte maturation, which encompasses dramatic potentiation of inositol 1,4,5-trisphosphate (IP3)-dependent Ca2+ release. Here we show that increased IP3 receptor (IP3R) sensitivity is initiated at the germinal vesicle breakdown stage of maturation, which correlates with maturation promoting factor (MPF) activation. Extensive phosphopeptide mapping of the IP3R resulted in ∼70% coverage and identified three residues, Thr-931, Thr-1136, and Ser-114, which are specifically phos pho ryl a ted during maturation. Phospho-specific antibody analyses show that Thr-1136 phos pho ryl a tion requires MPF activation. Activation of either MPF or the mitogen-activated protein kinase cascade independently, functionally sensitizes IP3-dependent Ca2+ release. Collectively, these data argue that the kinase cascades driving meiotic maturation potentiates IP3-dependent Ca2+ release, possibly trough direct phos pho ryl a tion of the IP3R.Egg activation refers to the cellular and molecular events that take place immediately following fertilization, transitioning the zygote into embryogenesis. In vertebrates, egg activation encompasses the block to polyspermy and the completion of oocyte meiosis, which is coupled to the extrusion of the second polar body. Interestingly, in all sexually reproducing organisms tested to date the cellular events associated with egg activation are Ca2+-dependent (1). Importantly the Ca2+ signal at fertilization encodes the progression of these cellular events in a defined temporal sequence that ensures a functional egg-to-embryo transition (2, 3). The first order of business for the fertilized egg is to block polyspermy, which could be lethal to the embryo. This presents a particularly difficult problem for the large Xenopus oocyte. Therefore, this species employs a fast and slow blocks to polyspermy, both of which are Ca2+-dependent (4). In addition, the Ca2+ release wave at fertilization releases the metaphase II cytostatic factor-dependent arrest in Xenopus oocytes. As is the case in other vertebrates, Xenopus eggs arrest at metaphase of meiosis II, an event that marks the completion of maturation.Therefore, Ca2+ dynamics at fertilization initiate and temporally encode critical cellular events for the egg-to-embryo transition. Specificity in Ca2+ signaling is encoded to a large extent in the spatial, temporal, and amplitude features of the Ca2+ signal. This endows Ca2+ signaling with its versatility and specificity, where in the same cell Ca2+ signals can mediate distinct cellular responses (5, 6).Ca2+ signaling pathways and intracellular organelles remodel during oocyte maturation, a complex cellular differentiation that prepares the egg for fertilization and egg activation (7, 8). In Xenopus the activity and distribution of multiple essential Ca2+-transporting proteins is modulated dramatically during oocyte maturation (8). Functional studies and mathematical modeling support the conclusion that the two critical determinants of Ca2+ signaling remodeling during Xenopus oocyte maturation are the internalization of the plasma-membrane Ca2+-ATPase, and the sensitization of inositol 1,4,5-trisphosphate (IP3)2-dependent Ca2+ release (911). Indeed Ca2+ release from intracellular stores through the IP3 receptor (IP3R) represents the primary source for the initial Ca2+ rise at fertilization in vertebrates (1214). The sensitivity of IP3-dependent Ca2+ release is enhanced during maturation (10, 15). The IP3R physically clusters during maturation (9, 16), and this is associated with functional clustering of elementary Ca2+ release events (10). IP3R clustering is important for the slow and continuous nature of Ca2+ wave propagation in Xenopus eggs (10). In fact the potentiation of IP3-dependent Ca2+ release is a hallmark of Ca2+ signaling differentiation during oocyte maturation in several vertebrate and invertebrate species (1719). However, the mechanisms underlying enhanced IP3-dependent Ca2+ release are not well understood.An attractive mechanism to explain increased IP3R sensitivity during oocyte maturation is phosphorylation, given the critical role kinase cascades play in the initiation and progression of the meiotic cell cycle. Furthermore, the affinity of the IP3R increases during mitosis apparently due to direct phosphorylation by maturation-promoting factor (MPF) (20, 21). In contrast, in starfish eggs, although the increase in Ca2+ release was dependent on MPF activation, MPF does not directly phosphorylate the IP3R, but rather it appears to mediate its effect through the actin cytoskeleton (22, 23). More recently, the MAPK cascade has been shown to be important for shaping Ca2+ dynamics in mouse eggs (24). Together, these results argue that phosphorylation plays an important role in the sensitization of IP3-dependent Ca2+ release during M-phase.Xenopus oocyte maturation is initiated by steroids that appear to act on a cell surface receptor (25). An important kinase cascade activated during maturation is the MAPK cascade that is initiated through the accumulation of Mos (Fig. 1A). This cascade culminates in the inhibition of Myt1, which phosphorylates and inhibits MPF. MPF is the key regulator of entry into M-phase and is composed of a Ser/Thr kinase subunit (cdk1) and cyclin B as a regulatory subunit. In addition, activation of Cdc25C is essential for oocyte maturation, because it represents the rate-limiting step in MPF activation (26). Cdc25C is phosphorylated by polo-like kinase through unknown upstream steps. In this work we analyze the functional regulation and phosphorylation pattern of the IP3R during oocyte maturation to better understand the role of cell cycle kinases in modulating IP3-dependent Ca2+ release.Open in a separate windowFIGURE 1.IP3-dependent Ca2+ release dynamics during maturation. A, kinase cascades driving Xenopus oocyte maturation. B, oocytes were injected with caged-IP3 and Oregan Green 1,2-bis(2-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid tetrakis 1 before imaging. Maturation was induced with progesterone, and cells were collected at different time points as indicated. Cells were imaged in line scan mode on a Zeiss LSM510 with the near UV 450 nm laser continuously on, at low intensity to produce a slow gradual IP3 rise. After imaging each cell was lysed and analyzed individually for the activation state of MAPK and MPF. MPF was assayed using an anti-phospho-Tyr-15-cdk1 antibody (arrow). Dephosphorylation is indicative of MPF activation. MAPK activation was detected using a phospho-specific MAPK antibody (arrowhead). Tubulin was the loading control (dash). C, percent of cells at each time point that either exhibit no release for the duration of the line scan (No Rel., black), puffs only (puffs, green), puffs followed by a wave (Puff-Wave, blue), or only a Ca2+ wave (Wave, red). For each time point n = 11–23 cells. D, amplitude of the first peak during the line scan as compared with the maximal Ca2+ signal. Mean ± S.E. (n = 9–18). E, latency until the first Ca2+ signal (Time to first peak) as compared with the time required to reach maximal signal (Time to Max). Mean ± S.E. (n = 9–18). For C–E: oocytes (Ooc); cells treated with progesterone that have not undergone GVBD at 2 or more hours after progesterone (p > 2); cells at GVBD and up to 0.5 h after GVBD (GVBD 0–0.5); cells from 0.5 to 2.5 h after GVBD (GVBD 0.5–2.5); fully mature eggs at 3 or more hours after GVBD (>3 egg).  相似文献   

4.
Ryanodine receptors (RyR) are Ca2+ channels that mediate Ca2+ release from intracellular stores in response to diverse intracellular signals. In RINm5F insulinoma cells, caffeine, and 4-chloro-m-cresol (4CmC), agonists of RyR, stimulated Ca2+ entry that was independent of store-operated Ca2+ entry, and blocked by prior incubation with a concentration of ryanodine that inactivates RyR. Patch-clamp recording identified small numbers of large-conductance (γK = 169 pS) cation channels that were activated by caffeine, 4CmC or low concentrations of ryanodine. Similar channels were detected in rat pancreatic β-cells. In RINm5F cells, the channels were blocked by cytosolic, but not extracellular, ruthenium red. Subcellular fractionation showed that type 3 IP3 receptors (IP3R3) were expressed predominantly in endoplasmic reticulum, whereas RyR2 were present also in plasma membrane fractions. Using RNAi selectively to reduce expression of RyR1, RyR2, or IP3R3, we showed that RyR2 mediates both the Ca2+ entry and the plasma membrane currents evoked by agonists of RyR. We conclude that small numbers of RyR2 are selectively expressed in the plasma membrane of RINm5F pancreatic β-cells, where they mediate Ca2+ entry.Ryanodine receptors (RyR)3 and inositol 1,4,5-trisphosphate receptors (IP3R) (1, 2) are the archetypal intracellular Ca2+ channels. Both are widely expressed, although RyR are more restricted in their expression than IP3R (3, 4). In common with many cells, pancreatic β-cells and insulin-secreting cell lines express both IP3R (predominantly IP3R3) (5, 6) and RyR (predominantly RyR2) (7). Both RyR and IP3R are expressed mostly within membranes of the endoplasmic (ER), where they mediate release of Ca2+. Functional RyR are also expressed in the secretory vesicles (8, 9) or, and perhaps more likely, in the endosomes of β-cells (10). Despite earlier suggestions (11), IP3R are probably not present in the secretory vesicles of β-cells (8, 12, 13).All three subtypes of IP3R are stimulated by IP3 with Ca2+ (1), and the three subtypes of RyR are each directly regulated by Ca2+. However, RyR differ in whether their most important physiological stimulus is depolarization of the plasma membrane (RyR1), Ca2+ (RyR2) or additional intracellular messengers like cyclic ADP-ribose. The latter stimulates both Ca2+ release and insulin secretion in β-cells (8, 14). The activities of both families of intracellular Ca2+ channels are also modulated by many additional signals that act directly or via phosphorylation (15, 16). Although they commonly mediate release of Ca2+ from the ER, both IP3R and RyR select rather poorly between Ca2+ and other cations (permeability ratio, PCa/PK ∼7) (1, 17). This may allow electrogenic Ca2+ release from the ER to be rapidly compensated by uptake of K+ (18), and where RyR or IP3R are expressed in other membranes it may allow them to affect membrane potential.Both Ca2+ entry and release of Ca2+ from intracellular stores contribute to the oscillatory increases in cytosolic Ca2+ concentration ([Ca2+]i) that stimulate exocytosis of insulin-containing vesicles in pancreatic β-cells (7). Glucose rapidly equilibrates across the plasma membrane (PM) of β-cells and its oxidative metabolism by mitochondria increases the cytosolic ATP/ADP ratio, causing KATP channels to close (19). This allows an unidentified leak current to depolarize the PM (20) and activate voltage-gated Ca2+ channels, predominantly L-type Ca2+ channels (21). The resulting Ca2+ entry is amplified by Ca2+-induced Ca2+ release from intracellular stores (7), triggering exocytotic release of insulin-containing dense-core vesicles (22). The importance of this sequence is clear from the widespread use of sulfonylurea drugs, which close KATP channels, in the treatment of type 2 diabetes. Ca2+ uptake by mitochondria beneath the PM further stimulates ATP production, amplifying the initial response to glucose and perhaps thereby contributing to the sustained phase of insulin release (23). However, neither the increase in [Ca2+]i nor the insulin release evoked by glucose or other nutrients is entirely dependent on Ca2+ entry (7, 24) or closure of KATP channels (25). This suggests that glucose metabolism may also more directly activate RyR (7, 26) and/or IP3R (27) to cause release of Ca2+ from intracellular stores. A change in the ATP/ADP ratio is one means whereby nutrient metabolism may be linked to opening of intracellular Ca2+ channels because both RyR (28) and IP3R (1) are stimulated by ATP.The other major physiological regulators of insulin release are the incretins: glucagon-like peptide-1 and glucose-dependent insulinotropic hormone (29). These hormones, released by cells in the small intestine, stimulate synthesis of cAMP in β-cells and thereby potentiate glucose-evoked insulin release (30). These pathways are also targets of drugs used successfully to treat type 2 diabetes (29). The responses of β-cells to cAMP involve both cAMP-dependent protein kinase and epacs (exchange factors activated by cAMP) (31, 32). The effects of the latter are, at least partly, due to release of Ca2+ from intracellular stores via RyR (3335) and perhaps also via IP3R (36). The interplays between Ca2+ and cAMP signaling generate oscillatory changes in the concentrations of both messengers (37). RyR and IP3R are thus implicated in mediating responses to each of the major physiological regulators of insulin secretion: glucose and incretins.Here we report that in addition to expression in intracellular stores, which probably include both the ER and secretory vesicles and/or endosomes, functional RyR2 are also expressed in small numbers in the PM of RINm5F insulinoma cells and rat pancreatic β-cells.  相似文献   

5.
ATP is known to increase the activity of the type-1 inositol 1,4,5-trisphosphate receptor (InsP3R1). This effect is attributed to the binding of ATP to glycine rich Walker A-type motifs present in the regulatory domain of the receptor. Only two such motifs are present in neuronal S2+ splice variant of InsP3R1 and are designated the ATPA and ATPB sites. The ATPA site is unique to InsP3R1, and the ATPB site is conserved among all three InsP3R isoforms. Despite the fact that both the ATPA and ATPB sites are known to bind ATP, the relative contribution of these two sites to the enhancing effects of ATP on InsP3R1 function is not known. We report here a mutational analysis of the ATPA and ATPB sites and conclude neither of these sites is required for ATP modulation of InsP3R1. ATP augmented InsP3-induced Ca2+ release from permeabilized cells expressing wild type and ATP-binding site-deficient InsP3R1. Similarly, ATP increased the single channel open probability of the mutated InsP3R1 to the same extent as wild type. ATP likely exerts its effects on InsP3R1 channel function via a novel and as yet unidentified mechanism.Inositol 1,4,5-trisphosphate receptors (InsP3R)3 are a family of large, tetrameric, InsP3-gated cation channels. The three members of this family (InsP3R1, InsP3R2, and InsP3R3) are nearly ubiquitously expressed and are localized primarily to the endoplasmic reticulum (ER) membrane (13). Numerous hormones, neurotransmitters, and growth factors bind to receptors that stimulate phospholipase C-induced InsP3 production (4). InsP3 subsequently binds to the InsP3R and induces channel opening. This pathway represents a major mechanism for Ca2+ liberation from ER stores (5). All three InsP3R isoforms are dynamically regulated by cytosolic factors in addition to InsP3 (1). Ca2+ is perhaps the most important determinant of InsP3R activity besides InsP3 itself and is known to regulate InsP3R both positively and negatively (6). ATP, in concert with InsP3 and Ca2+, also regulates InsP3R as do numerous kinases, phosphatases, and protein-binding partners (710). This intricate network of regulation allows InsP3R activity to be finely tuned by the local cytosolic environment (9). As a result, InsP3-induced Ca2+ signals can exhibit a wide variety of spatial and temporal patterns, which likely allows Ca2+ to control many diverse cellular processes.Modulation of InsP3-induced Ca2+ release (IICR) by ATP and other nucleotides provides a direct link between intracellular Ca2+ signaling and the metabolic state of the cell. Metabolic fluctuations could, therefore, impact Ca2+ signaling in many cell types given that InsP3R are expressed in all cells (11, 12). Consistent with this, ATP has been shown to augment IICR in many diverse cell types including primary neurons (13), smooth muscle cells (14), and exocrine acinar cells (15) as well as in immortalized cell lines (1618). The effects of ATP on InsP3R function do not require hydrolysis because non-hydrolyzable ATP analogues are as effective as ATP (7, 14). ATP is thought to bind to distinct regions in the central, coupling domain of the receptors and to facilitate channel opening (2, 19). ATP is not required for channel gating, but instead, increases InsP3R activity in an allosteric fashion by increasing the open probability of the channel in the presence of activating concentrations of InsP3 and Ca2+ (7, 8, 20).Despite a wealth of knowledge regarding the functional effects of ATP on InsP3R function, there is relatively little known about the molecular determinants of these actions. ATP is thought to exert effects on channel function by direct binding to glycine-rich regions containing the consensus sequence GXGXXG that are present in the receptors (2). These sequences were first proposed to be ATP-binding domains due to their similarity with Walker A motifs (21). The neuronal S2+ splice variant of InsP3R1 contains two such domains termed ATPA and ATPB. A third site, ATPC, is formed upon removal of the S2 splice site (2, 22). The ATPB site is conserved in InsP3R2 and InsP3R3, while the ATPA and ATPC sites are unique to InsP3R1. Our prior work examining the functional consequences of mutating these ATP-binding sites has yielded unexpected results. For example, mutating the ATPB site in InsP3R2 completely eliminated the enhancing effects of ATP on this isoform while mutating the analogous site in InsP3R3 failed to alter the effects of ATP (23). This indicated the presence of an additional locus for ATP modulation of InsP3R3. In addition, mutation of the ATPC in the S2 splice variant of InsP3R1 did not alter the ability of ATP to modulate Ca2+ release, but instead impaired the ability of protein kinase A to phosphorylate Ser-1755 of this isoform (22).The ATPA and ATPB sites in InsP3R1 were first identified as putative nucleotide-binding domains after the cloning of the full-length receptor (24). Early binding experiments with 8-azido-[α-32P]ATP established that ATP cross-linked with receptor purified from rat cerebellum at one site per receptor monomer (19). Later, more detailed, binding experiments on trypsinized recombinant rat InsP3R1 showed cross-linking of ATP to two distinct regions of the receptor that corresponded with the ATPA and ATPB sites (17). We and others (16, 22, 23) have also reported the binding of ATP analogues to purified GST fusions of small regions of InsP3R1 surrounding the ATPA and ATPB sites. It is widely accepted, in the context of the sequence similarity to Walker A motifs and biochemical data, that the ATPA and ATPB sites are the loci where ATP exerts its positive functional effects on InsP3R1 function (13, 16). Furthermore, the higher affinity of the ATPA site to ATP is thought to confer the higher sensitivity of InsP3R1 to ATP versus InsP3R3, which contains the ATPB site exclusively (25, 26). The purpose of this study, therefore, was to examine the contributions of the ATPA and ATPB sites to ATP modulation of the S2+ splice variant of InsP3R1. We compared the effects of ATP on InsP3R1 and on ATP-binding site mutated InsP3R1 using detailed functional analyses in permeabilized cells and in single channel recordings. Here we report that InsP3R1 is similar to InsP3R3 in that ATP modulates IICR even at maximal InsP3 concentrations and that neither the ATPA nor the ATPB site is required for this effect.  相似文献   

6.
Protein kinase A (PKA) phosphorylation of inositol 1,4,5-trisphosphate receptors (InsP3Rs) represents a mechanism for shaping intracellular Ca2+ signals following a concomitant elevation in cAMP. Activation of PKA results in enhanced Ca2+ release in cells that express predominantly InsP3R2. PKA is known to phosphorylate InsP3R2, but the molecular determinants of this effect are not known. We have expressed mouse InsP3R2 in DT40-3KO cells that are devoid of endogenous InsP3R and examined the effects of PKA phosphorylation on this isoform in unambiguous isolation. Activation of PKA increased Ca2+ signals and augmented the single channel open probability of InsP3R2. A PKA phosphorylation site unique to the InsP3R2 was identified at Ser937. The enhancing effects of PKA activation on this isoform required the phosphorylation of Ser937, since replacing this residue with alanine eliminated the positive effects of PKA activation. These results provide a mechanism responsible for the enhanced Ca2+ signaling following PKA activation in cells that express predominantly InsP3R2.Hormones, neurotransmitters, and growth factors stimulate the production of InsP33 and Ca2+ signals in virtually all cell types (1). The ubiquitous nature of this mode of signaling dictates that this pathway does not exist in isolation; indeed, a multitude of additional signaling pathways can be activated simultaneously. A prime example of this type of “cross-talk” between independently activated signaling systems results from the parallel activation of cAMP and Ca2+ signaling pathways (2, 3). Interactions between these two systems occur in numerous distinct cell types with various physiological consequences (36). Given the central role of InsP3R in Ca2+ signaling, a major route of modulating the spatial and temporal features of Ca2+ signals following cAMP production is potentially through PKA phosphorylation of the InsP3R isoform(s) expressed in a particular cell type.There are three InsP3R isoforms (InsP3R1, InsP3R2, and InsP3R3) expressed to varying degrees in mammalian cells (7, 8). InsP3R1 is the major isoform expressed in the nervous system, but it is less abundant compared with other subtypes in non-neuronal tissues (8). Ca2+ release via InsP3R2 and InsP3R3 predominate in these tissues. InsP3R2 is the major InsP3R isoform in many cell types, including hepatocytes (7, 8), astrocytes (9, 10), cardiac myocytes (11), and exocrine acinar cells (8, 12). Activation of PKA has been demonstrated to enhance InsP3-induced Ca2+ signaling in hepatocytes (13) and parotid acinar cells (4, 14). Although PKA phosphorylation of InsP3R2 is a likely causal mechanism underlying these effects, the functional effects of phosphorylation have not been determined in cells unambiguously expressing InsP3R2 in isolation. Furthermore, the molecular determinants of PKA phosphorylation of this isoform are not known.PKA-mediated phosphorylation is an efficient means of transiently and reversibly regulating the activity of the InsP3R. InsP3R1 was identified as a major substrate of PKA in the brain prior to its identification as the InsP3R (15, 16). However, until recently, the functional consequences of phosphorylation were unresolved. Initial conflicting results were reported indicating that phosphoregulation of InsP3R1 could result in either inhibition or stimulation of receptor activity (16, 17). Mutagenic strategies were employed by our laboratory to clarify this discrepancy. These studies unequivocally assigned phosphorylation-dependent enhanced Ca2+ release and InsP3R1 activity at the single channel level, through phosphorylation at canonical PKA consensus motifs at Ser1589 and Ser1755. The sites responsible were also shown to be specific to the particular InsP3R1 splice variant (18). These data were also corroborated by replacing the relevant serines with glutamates in a strategy designed to construct “phosphomimetic” InsP3R1 by mimicking the negative charge added by phosphorylation (19, 20). Of particular note, however, although all three isoforms are substrates for PKA, neither of the sites phosphorylated by PKA in InsP3R1 are conserved in the other two isoforms (21). Recently, three distinct PKA phosphorylation sites were identified in InsP3R3 that were in different regions of the protein when compared with InsP3R1 (22). To date, no PKA phosphorylation sites have been identified in InsP3R2.Interactions between Ca2+ and cAMP signaling pathways are evident in exocrine acinar cells of the parotid salivary gland. In these cells, both signals are important mediators of fluid and protein secretion (23). Multiple components of the [Ca2+]i signaling pathway in these cells are potential substrates for modulation by PKA. Previous work from this laboratory established that activation of PKA potentiates muscarinic acetylcholine receptor-induced [Ca2+]i signaling in mouse and human parotid acinar cells (4, 24, 25). A likely mechanism to explain this effect is that PKA phosphorylation increases the activity of InsP3R expressed in these cells. Consistent with this idea, activation of PKA enhanced InsP3-induced Ca2+ release in permeabilized mouse parotid acinar cells and also resulted in the phosphorylation of InsP3R2 (4).Invariably, prior work examining the functional effects of PKA phosphorylation on InsP3R2 has been performed using cell types expressing multiple InsP3R isoforms. For example, AR4-2J cells are the preferred cell type for examining InsP3R2 in relative isolation, because this isoform constitutes more than 85% of the total InsP3R population (8). InsP3R1, however, contributes up to ∼12% of the total InsP3R in AR4-2J cells. An initial report using InsP3-mediated 45Ca2+ flux suggested that PKA activation increased InsP3R activity in AR4-2J cells (21). A similar conclusion was made in a later study, which documented the effects of PKA activation on agonist stimulated Ca2+ signals in AR4-2J cells (26). Any effects of phosphorylation observed in these experiments could plausibly have resulted from phosphorylation of the residual InsP3R1.Although PKA enhances InsP3-induced calcium release in cells expressing predominantly InsP3R2, including hepatocytes, parotid acinar cells, and AR4-2J cells (4, 13, 21, 26, 27), InsP3R2 is not phosphorylated at stoichiometric levels by PKA (21). This observation has called into question the physiological significance of PKA phosphorylation of InsP3R2 (28). The apparent low levels of InsP3R2 phosphorylation are clearly at odds with the augmented Ca2+ release observed in cells expressing predominantly this isoform. The equivocal nature of these findings probably stems from the fact that, to date, all of the studies demonstrating positive effects of PKA activation on Ca2+ release were conducted in cells that also express InsP3R1. The purpose of the current experiments was to analyze the functional effects of phosphorylation on InsP3R2 expressed in isolation on a null background. We report that InsP3R2 activity is increased by PKA phosphorylation under these conditions, and furthermore, we have identified a unique phosphorylation site in InsP3R2 at Ser937. In total, these results provide a direct mechanism for the cAMP-induced activation of InsP3R2 via PKA phosphorylation of InsP3R2.  相似文献   

7.
Intracellular Ca2+ mobilization plays an important role in a wide variety of cellular processes, and multiple second messengers are responsible for mediating intracellular Ca2+ changes. Here we explored the role of one endogenous Ca2+-mobilizing nucleotide, cyclic adenosine diphosphoribose (cADPR), in the proliferation and differentiation of neurosecretory PC12 cells. We found that cADPR induced Ca2+ release in PC12 cells and that CD38 is the main ADP-ribosyl cyclase responsible for the acetylcholine (ACh)-induced cADPR production in PC12 cells. In addition, the CD38/cADPR signaling pathway is shown to be required for the ACh-induced Ca2+ increase and cell proliferation. Inhibition of the pathway, on the other hand, accelerated nerve growth factor (NGF)-induced neuronal differentiation in PC12 cells. Conversely, overexpression of CD38 increased cell proliferation but delayed NGF-induced differentiation. Our data indicate that cADPR plays a dichotomic role in regulating proliferation and neuronal differentiation of PC12 cells.Mobilization of intracellular Ca2+ stores is involved in diverse cell functions, including fertilization, cell proliferation, and differentiation (14). At least three endogenous Ca2+-mobilizing messengers have been identified, including inositol trisphosphate (IP3),3 nicotinic adenine acid dinucleotide phosphate (NAADP), and cyclic adenosine diphosphoribose (cADPR). Similar to IP3, cADPR can mobilize calcium release in a wide variety of cell types and species, from protozoa to animals. The cADPR-mediated Ca2+ signaling has been indicated in a variety of cellular processes (57), from abscisic acid signaling and regulation of the circadian clock in plants, to mediating long-term synaptic depression in hippocampus.Ample evidence shows that the ryanodine receptors are the main intracellular targets for cADPR (1, 2, 8). Ryanodine receptors (RyRs) are intracellular Ca2+ channels widely expressed in various cells and tissues, including muscles and neurons. It is the major cellular mediator of Ca2+-induced Ca2+ release (CICR) in cells. There are three isoforms of ryanodine receptors: RyR1, RyR2, and RyR3, all of which have been implicated in the cADPR signaling (1, 2, 8). However, evidence regarding cADPR acting directly on the receptors is lacking (9). It has been suggested that accessory proteins, such as calmodulin and FK506-binding protein (FKBP), may be involved instead (1015).cADPR is formed from nicotinamide adenine dinucleotide (NAD) by ADP-ribosyl cyclases. Six ADP-ribosyl cyclases have been identified so far: Aplysia ADP-ribosyl cyclase, three sea urchin homologues (16, 17), and two mammalian homologues, CD38 and CD157 (18). CD38 is a membrane-bound protein and the main mammalian ADP-ribosyl cyclase. As a novel multifunctional enzyme, CD38 catalyzes the synthesis and hydrolysis of both cADPR and NAADP, two structurally and functionally distinct Ca2+ messengers. Virtually all mammalian tissues ever examined have been shown to express CD38. CD38 knock-out mice exhibit multiple physiological defects, ranging from impaired immune responses, metabolic disturbances, to behavioral modifications (1, 6, 18).CD38 was originally identified as a lymphocyte differentiation antigen (18). Indeed, CD38/cADPR has been linked to cell differentiation (5). For example, in human HL-60 cells, CD38 expression and the consequential accumulation of cADPR play a causal role in mediating granulocytic differentiation (19). In addition, expression of CD38 in HeLa and 3T3 cells not only increased intracellular Ca2+ concentration but also induced cell proliferation by significantly reducing the S phase duration, leading to shortened cell doubling time (20). The ability of cADPR to increase cell proliferation has also been observed in human T cells (21), human hemopoietic progenitors (22), human peripheral blood mononuclear cells (23), human mesenchymal stem cells (24), and murine mesangial cells (25).The PC12 cell line was derived from rat adrenal medulla and has been used extensively as a neuronal model, since it exhibits many of the functions observed in primary neuronal cultures (26). Most importantly, PC12 cells can be induced by nerve growth factor (NGF) to differentiate into cells with extensive neurite outgrowths, resembling neuronal dendritic trees (26, 27). In contrast to NGF, numerous growth factors and neurotransmitters can induce the proliferation of PC12 cells instead (26). Both IP3 receptor- and ryanodine receptor-mediated Ca2+ stores have been shown to be present in PC12 cells (2831). The type 2 ryanodine receptor is expressed in PC12 cells and activation of the NO/cGMP pathway in PC12 cells results in calcium mobilization, which is mediated by cADPR and similar to that seen in sea urchin eggs (32). It has been demonstrated that NAADP, another Ca2+-mobilizing messenger, is also a potent neuronal differentiation inducer in PC12 cells, while IP3 exhibits no such role (33, 34). Whether cADPR is involved in the proliferation and differentiation of PC12 cells is unknown.Here we show that activation of the CD38/cADPR/Ca2+ signaling is required for the ACh-induced proliferation in PC12 cells, while inhibition of the pathway accelerates NGF-induced neuronal differentiation. Our data indicate that cADPR is important in regulating cell proliferation and neuronal differentiation in PC12 cells.  相似文献   

8.
A database search of the Paramecium genome reveals 34 genes related to Ca2+-release channels of the inositol-1,4,5-trisphosphate (IP3) or ryanodine receptor type (IP3R, RyR). Phylogenetic analyses show that these Ca2+ release channels (CRCs) can be subdivided into six groups (Paramecium tetraurelia CRC-I to CRC-VI), each one with features in part reminiscent of IP3Rs and RyRs. We characterize here the P. tetraurelia CRC-IV-1 gene family, whose relationship to IP3Rs and RyRs is restricted to their C-terminal channel domain. CRC-IV-1 channels localize to cortical Ca2+ stores (alveolar sacs) and also to the endoplasmic reticulum. This is in contrast to a recently described true IP3 channel, a group II member (P. tetraurelia IP3RN-1), found associated with the contractile vacuole system. Silencing of either one of these CRCs results in reduced exocytosis of dense core vesicles (trichocysts), although for different reasons. Knockdown of P. tetraurelia IP3RN affects trichocyst biogenesis, while CRC-IV-1 channels are involved in signal transduction since silenced cells show an impaired release of Ca2+ from cortical stores in response to exocytotic stimuli. Our discovery of a range of CRCs in Paramecium indicates that protozoans already have evolved multiple ways for the use of Ca2+ as signaling molecule.Ca2+ is an important component of cell activity in all organisms, from protozoa to mammals. Thereby Ca2+ may originate from the outside medium and/or from internal stores (7, 18). Ca2+ release from internal stores is mediated by various Ca2+ release channels (CRCs), of which the inositol-1,4,5-trisphosphate receptor (IP3R) and ryanodine receptor (RyR) families have been studied most extensively (8, 9, 29, 63). IP3Rs and RyRs have been identified in various metazoan organisms (reviewed in references 9, 28, and 104). According to these reviews, there exist three genetically distinct isoforms of each receptor type in mammals and orthologues have been identified in various nonmammalian vertebrates, e.g., frogs, chickens, and fish. RyRs and IP3Rs were also cloned and sequenced in the invertebrates Drosophila melanogaster and Caenorhabditis elegans, which possess one copy of each receptor type.Functional evidence for Ca2+ release in response to ryanodine or IP3 receptor agonists has been described in several unicellular systems. Treatment of permeabilized Plasmodium chabaudi parasites with IP3 results in Ca2+ release, which is inhibited by the IP3 receptor antagonist heparin (69). Another apicomplexan parasite, Toxoplasma gondii, responds to agonists and antagonists of both, ryanodine and IP3 receptors, by mediating increases in intracellular Ca2+ concentration ([Ca2+]i) (56). Stimulation of Trypanosoma cruzi with carbachol results in increased [Ca2+]i and IP3 (59). IP3 and cyclic ADP-ribose induces Ca2+ release in Euglena gracilis microsome fractions in a dose-dependent manner (61). In the giant algae Chara corallina and Nitrella translucens, IP3 produces action potentials involving increased [Ca2+]i (93). Treatment of vacuolar membrane vesicles from Candida albicans with IP3 results in Ca2+ release, blocked by heparin and ruthenium red (14). IP3 generates and maintains a Ca2+ gradient in the hyphal tip of Neurospora crassa and the IP3-sensitive channels have been reconstituted and characterized with the planar bilayer method (87). In summary, these publications suggest that IP3-dependent signaling pathways are conserved among unicellular organisms, including protozoa.Despite these data, the molecular characterization of IP3 or ryanodine receptors in low eukaryotes is currently a challenge since the identification of orthologues has not been possible thus far, probably because of evolutionary sequence divergence (66). Traynor et al. (96) identified an IP3 receptor-like protein, IplA, in Dictyostelium discoideum, which possesses regions related to IP3R sequences, but thus far no evidence for IP3 interaction exists. We have recently described an IP3R in the ciliated protozoa Paramecium tetraurelia (referred to here as P. tetraurelia IP3RN) (53), with features characteristic of mammalian IP3Rs in terms of topology and ability for IP3 binding. The expression level of P. tetraurelia IP3RN is modulated by extracellular Ca2+ concentrations ([Ca2+]o) and immunofluorescence studies reveal an unexpected localization to the contractile vacuole complex (CVC), the major organelle involved in osmoregulation (2). The ionic composition of the contractile vacuole fluid by ion-selective microelectrodes (91) suggests that the organelle plays a major role in expelling an excess of cytosolic Ca2+. Therefore, these IP3Rs may here mediate a latent, graded reflux of Ca2+ for fine-tuning of [Ca2+]i and thus serve [Ca2+] homeostasis (53).Besides [Ca2+] homeostasis, the Paramecium cell has to regulate a variety of well-characterized processes (75). This includes exocytosis of dense-core secretory vesicles (trichocysts) (71, 74, 99). Each cell possesses up to 1,000 trichocysts attached to the cell membrane. Their contents can be extruded synchronously in response to natural stimuli, i.e., predators (34, as confirmed by Knoll et al. [49]), to artificial polyamine secretagogues such as aminoethyldextran (AED) (78), to caffeine (48) or to the ryanodine substitute, 4-chloro-meta-cresol (4-CmC) (46). Their expulsion strictly depends on Ca2+ (10) and is accompanied by an increase of intracellular [Ca2+]i (24, 47). This Ca2+ signal originates from rapid mobilization of cortical stores, the alveolar sacs (33, 64, 74), superimposed by Ca2+ influx (46, 72). It thus represents a SOC-type mechanism (SOC, store-operated Ca2+ entry) known from mammalian systems (81).Upon exocytosis stimulation ∼60% of their total Ca2+ is released from alveolar sacs (33). These are Ca2+ stores (90) represented by flat membrane compartments tightly attached at the cell membrane surrounding each trichocyst docking site. They possess a SERCA-type pump located at the membrane facing the cell center (36, 37) and a luminal high-capacity/low-affinity CaBP of the calsequestrin type (73). Thus far, Ca2+ release channels of these stores were identified only indirectly as cells respond by exocytosis to the RyR activators caffeine (54, 48) and 4-CmC (46). However, an involvement of conserved RyRs has remained questionable as ryanodine is not able to activate Ca2+ release from alveolar sacs, as is the case with IP3 (54). Therefore, one of the most intriguing questions is the elucidation of the molecular nature of the channels mediating Ca2+ release from alveolar sacs upon stimulated exocytosis.In the present work we describe a novel family of CRCs (P. tetraurelia CRC-IV-1), whose members display several properties of the channels postulated above. In detail, the identified CRC-IV-1 channels localize to the alveolar sacs. Functional and fluorochrome analyses after gene silencing reveal that they are essential for mediating Ca2+ release and exocytosis in response to AED, caffeine, or 4-CmC. Their classification as “novel” CRC type is based on a restricted relationship to the C-terminal channel domains of IP3Rs and RyRs. The overall size and the number of putative transmembrane domains resemble IP3Rs, but N-terminal parts of CRC-IV-1 channels do not show any conservation, such as an IP3-binding domain. Therefore, CRC-IV-1 channels represent distant relatives of IP3Rs and RyRs and may belong to an ancestral Ca2+ signaling pathway.  相似文献   

9.
Calmodulin binds to IQ motifs in the α1 subunit of CaV1.1 and CaV1.2, but the affinities of calmodulin for the motif and for Ca2+ are higher when bound to CaV1.2 IQ. The CaV1.1 IQ and CaV1.2 IQ sequences differ by four amino acids. We determined the structure of calmodulin bound to CaV1.1 IQ and compared it with that of calmodulin bound to CaV1.2 IQ. Four methionines in Ca2+-calmodulin form a hydrophobic binding pocket for the peptide, but only one of the four nonconserved amino acids (His-1532 of CaV1.1 and Tyr-1675 of CaV1.2) contacts this calmodulin pocket. However, Tyr-1675 in CaV1.2 contributes only modestly to the higher affinity of this peptide for calmodulin; the other three amino acids in CaV1.2 contribute significantly to the difference in the Ca2+ affinity of the bound calmodulin despite having no direct contact with calmodulin. Those residues appear to allow an interaction with calmodulin with one lobe Ca2+-bound and one lobe Ca2+-free. Our data also provide evidence for lobe-lobe interactions in calmodulin bound to CaV1.2.The complexity of eukaryotic Ca2+ signaling arises from the ability of cells to respond differently to Ca2+ signals that vary in amplitude, duration, and location. A variety of mechanisms decode these signals to drive the appropriate physiological responses. The Ca2+ sensor for many of these physiological responses is the Ca2+-binding protein calmodulin (CaM).2 The primary sequence of CaM is tightly conserved in all eukaryotes, yet it binds and regulates a broad set of target proteins in response to Ca2+ binding. CaM has two domains that bind Ca2+ as follows: an amino-terminal domain (N-lobe) and a carboxyl-terminal domain (C-lobe) joined via a flexible α-helix. Each lobe of CaM binds two Ca2+ ions, and binding within each lobe is highly cooperative. The two lobes of CaM, however, have distinct Ca2+ binding properties; the C-lobe has higher Ca2+ affinity because of a slower rate of dissociation, whereas the N-lobe has weaker Ca2+ affinity and faster kinetics (1). CaM can also bind to some target proteins in both the presence and absence of Ca2+, and the preassociation of CaM in low Ca2+ modulates the apparent Ca2+ affinity of both the amino-terminal and carboxyl-terminal lobes. Differences in the Ca2+ binding properties of the lobes and in the interaction sites of the amino- and carboxyl-terminal lobes enable CaM to decode local versus global Ca2+ signals (2).Even though CaM is highly conserved, CaM target (or recognition) sites are quite heterogeneous. The ability of CaM to bind to very different targets is at least partially due to its flexibility, which allows it to assume different conformations when bound to different targets. CaM also binds to various targets in distinct Ca2+ saturation states as follows: Ca2+-free (3), Ca2+ bound to only one of the two lobes, or fully Ca2+-bound (47). In addition, CaM may bind with both lobes bound to a target (5, 6) or with only a single lobe engaged (8). If a target site can bind multiple conformers of CaM, CaM may undergo several transitions that depend on Ca2+ concentration, thereby tuning the functional response. Identification of stable intermediate states of CaM bound to individual targets will help to elucidate the steps involved in this fine-tuned control.Both CaV1.1 and CaV1.2 belong to the L-type family of voltage-dependent Ca2+ channels, which bind apoCaM and Ca2+-CaM at carboxyl-terminal recognition sites in their α1 subunits (914). Ca2+ binding to CaM, bound to CaV1.2 produces Ca2+-dependent facilitation (CDF) (14). Whether CaV1.1 undergoes CDF is not known. However, both CaV1.2 and CaV1.1 undergo Ca2+- and CaM-dependent inactivation (CDI) (14, 15). CaV1.1 CDI is slower and more sensitive to buffering by 1,2-bis(o-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid than CaV1.2 CDI (15). Ca2+ buffers are thought to influence CDI and/or CDF in voltage-dependent Ca2+ channels by competing with CaM for Ca2+ (16).The conformation of the carboxyl terminus of the α1 subunit is critical for channel function and has been proposed to regulate the gating machinery of the channel (17, 18). Several interactions of this region include intramolecular contacts with the pore inactivation machinery and intermolecular contacts with CaM kinase II and ryanodine receptors (17, 1922). Ca2+ regulation of CaV1.2 may involve several motifs within this highly conserved region, including an EF hand motif and three contiguous CaM-binding sequences (10, 12). ApoCaM and Ca2+-CaM-binding sites appear to overlap at the site designated as the “IQ motif” (9, 12, 13), which are critical for channel function at the molecular and cellular level (14, 23).Differences in the rate at which 1,2-bis(o-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid affects CDI of CaV1.1 and CaV1.2 could reflect differences in their interactions with CaM. In this study we describe the differences in CaM interactions with the IQ motifs of the CaV1.1 and the CaV1.2 channels in terms of crystal structure, CaM affinity, and Ca2+ binding to CaM. We find the structures of Ca2+-CaM-IQ complexes are similar except for a single amino acid change in the peptide that contributes to its affinity for CaM. We also find that the other three amino acids that differ in CaV1.2 and CaV1.1 contribute to the ability of CaV1.2 to bind a partially Ca2+-saturated form of CaM.  相似文献   

10.
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   

11.
12.
13.
STIM1 and ORAI1, the two limiting components in the Ca2+ release-activated Ca2+ (CRAC) signaling cascade, have been reported to interact upon store depletion, culminating in CRAC current activation. We have recently identified a modulatory domain between amino acids 474 and 485 in the cytosolic part of STIM1 that comprises 7 negatively charged residues. A STIM1 C-terminal fragment lacking this domain exhibits enhanced interaction with ORAI1 and 2–3-fold higher ORAI1/CRAC current densities. Here we focused on the role of this CRAC modulatory domain (CMD) in the fast inactivation of ORAI1/CRAC channels, utilizing the whole-cell patch clamp technique. STIM1 mutants either with C-terminal deletions including CMD or with 7 alanines replacing the negative amino acids within CMD gave rise to ORAI1 currents that displayed significantly reduced or even abolished inactivation when compared with STIM1 mutants with preserved CMD. Consistent results were obtained with cytosolic C-terminal fragments of STIM1, both in ORAI1-expressing HEK 293 cells and in RBL-2H3 mast cells containing endogenous CRAC channels. Inactivation of the latter, however, was much more pronounced than that of ORAI1. The extent of inactivation of ORAI3 channels, which is also considerably more prominent than that of ORAI1, was also substantially reduced by co-expression of STIM1 constructs missing CMD. Regarding the dependence of inactivation on Ca2+, a decrease in intracellular Ca2+ chelator concentrations promoted ORAI1 current fast inactivation, whereas Ba2+ substitution for extracellular Ca2+ completely abrogated it. In summary, CMD within the STIM1 cytosolic part provides a negative feedback signal to Ca2+ entry by triggering fast Ca2+-dependent inactivation of ORAI/CRAC channels.The Ca2+ release-activated Ca2+ (CRAC)5 channel is one of the best characterized store-operated entry pathways (17). Substantial efforts have led to identification of two key components of the CRAC channel machinery: the stromal interaction molecule 1 (STIM1), which is located in the endoplasmic reticulum and acts as a Ca2+ sensor (810), and ORAI1/CRACM1, the pore-forming subunit of the CRAC channel (1113). Besides ORAI1, two further homologues named ORAI2 and ORAI3 belong to the ORAI channel family (12, 14).STIM1 senses endoplasmic reticulum store depletion primarily by its luminal EF-hand in its N terminus (8, 15), redistributes close to the plasma membrane, where it forms puncta-like structures, and co-clusters with ORAI1, leading to inward Ca2+ currents (12, 1619). The STIM1 C terminus, located in the cytosol, contains two coiled-coil regions overlapping with an ezrin-radixin-moesin (ERM)-like domain followed by a serine/proline- and a lysine-rich region (2, 8, 2022). Three recent studies have described the essential ORAI-activating region within the ERM domain, termed SOAR (Stim ORAI-activating region) (23), OASF (ORAI-activating small fragment) (24), and CAD (CRAC-activating domain) (25), including the second coiled coil domain and the following ∼55 amino acids. We and others have provided evidence that store depletion leads to a dynamic coupling of STIM1 to ORAI1 (2628) that is mediated by a direct interaction of the STIM1 C terminus with ORAI1 C terminus probably involving the putative coiled-coil domain in the latter (27).Furthermore, different groups have proven that the C terminus of STIM1 is sufficient to activate CRAC as well as ORAI1 channels independent of store depletion (2225, 27, 29). We have identified that OASF-(233–474) or shorter fragments exhibit further enhanced coupling to ORAI1 resulting in 3-fold increased constitutive Ca2+ currents. A STIM1 fragment containing an additional cluster of anionic amino acids C-terminal to position 474 displays weaker interaction with ORAI1 as well as reduced Ca2+ current comparable with that mediated by wild-type STIM1 C terminus. Hence, we have suggested that these 11 amino acids (474–485) act in a modulatory manner onto ORAI1; however, their detailed mechanistic impact within the STIM1/ORAI1 signaling machinery has remained so far unclear.In this study, we focused on the impact of this negative cluster on fast inactivation of STIM1-mediated ORAI Ca2+ currents. Lis et al. (30) have shown that all three ORAI homologues display distinct inactivation profiles, where ORAI2 and ORAI3 show a much more pronounced fast inactivation than ORAI1. Moreover, it has been reported (31) that different expression levels of STIM1 to ORAI1 affect the properties of CRAC current inactivation. Yamashita et al. (32) have demonstrated a linkage between the selectivity filter of ORAI1 and its Ca2+-dependent fast inactivation. Here we provide evidence that a cluster of acidic residues within the C terminus of STIM1 is involved in the fast inactivation of ORAI1 and further promotes that of ORAI3 and native CRAC currents.  相似文献   

14.
By now, little is known on L-type calcium channel (LTCC) subunits expressed in mouse heart. We show that CaVβ2 proteins are the major CaVβ components of the LTCC in embryonic and adult mouse heart, but that in embryonic heart CaVβ3 proteins are also detectable. At least two CaVβ2 variants of ∼68 and ∼72 kDa are expressed. To identify the underlying CaVβ2 variants, cDNA libraries were constructed from poly(A)+ RNA isolated from hearts of 7-day-old and adult mice. Screening identified 60 independent CaVβ2 cDNA clones coding for four types of CaVβ2 proteins only differing in their 5′ sequences. CaVβ2-N1, -N4, and -N5 but not -N3 were identified in isolated cardiomyocytes by RT-PCR and were sufficient to reconstitute the CaVβ2 protein pattern in vitro. Significant L-type Ca2+ currents (ICa) were recorded in HEK293 cells after co-expression of CaV1.2 and CaVβ2. Current kinetics were determined by the type of CaVβ2 protein, with the ∼72-kDa CaVβ2a-N1 shifting the activation of ICa significantly to depolarizing potentials compared with the other CaVβ2 variants. Inactivation of ICa was accelerated by CaVβ2a-N1 and -N4, which also lead to slower activation compared with CaVβ2a-N3 and -N5. In summary, this study reveals the molecular LTCC composition in mouse heart and indicates that expression of various CaVβ2 proteins may be used to adapt the properties of LTCCs to changing myocardial requirements during development and that CaVβ2a-N1-induced changes of ICa kinetics might be essential in embryonic heart.Cardiac contractions require Ca2+ influx in cardiomyocytes from the extracellular fluid, which leads to Ca2+ release from the sarcoplasmic reticulum via ryanodine receptors (1).This Ca2+-induced Ca2+ release (CICR)4 causes a marked increase in intracellular Ca2+ concentration for short periods of time and underlies cardiac contraction (2, 3). The Ca2+ influx into cardiac myocytes is mediated by high voltage-activated L-type Ca2+ channels (LTCCs), which are heteromultimeric complexes comprised predominantly of the pore-forming CaVα1 subunit and the auxiliary CaVβ subunit (4). In heart, the principal CaVα1 subunit, CaVα1c (CaV1.2), is encoded by the Cacna1C gene (5). Four genes (Cacnb1-4) encoding CaVβ subunits have been identified that are expressed in the heart of different species including human, rabbit, and rat (6, 7, 8).CaVβ proteins are ∼500 amino acid cytoplasmic proteins that bind to the CaVα1 I-II intracellular loop (9) and affect channel gating properties (4), trafficking (10, 11), regulation by neurotransmitter receptors through G-protein βγ subunit activation (12), and sensitivity to drugs (13). The CaVβ primary sequence encodes five domains, arranged V1-C1-V2-C2-V3. V1, V2, and V3 are variable domains, whereas C1 and C2 are conserved (14). Structural studies reveal that C1 and C2 form a SH3 domain (Src homology 3 domain) and a NK domain (nucleotide kinase domain), respectively (15). Although C1-V2-C2 makes the CaVβ core, in heart the V1 region appears critical for the kinetics of ICa and heart function. Accordingly a mutation in the V1 region of the Cacnb2 gene was recently identified as an underlying cause of Brugada syndrome (16).In mice-targeted deletion of the Cacnb2 gene (17) but not of Cacnb1 (18), Cacnb3 (19, 20), or Cacnb4 (21) leads to a morphologically and functionally compromised heart, which causes severe defective remodeling of intra- and extra-embryonic blood vessels and death at early embryonic stages both when the Cacnb2 gene was targeted globally or in a cardiac myocyte-specific way (17). Although these results point to an essential role of CaVβ2 for ICa and cardiac function, the existence of various CaVβ2 splice variants and heterogeneity of the expressed CaVβ2 proteins require further studies on the subunit composition of LTCCs in the mouse heart. In addition and in view of the growing number of preclinical studies using mouse models carrying definite Ca2+ channel subunits as transgenes in heart tissue, the identification of the relevant gene products underlying the endogenous mouse cardiac L-type channel is essential. Recent mouse models (e.g. 22, 23, 24) carrying a rat CaVβ2 splice variant (“rat CaVβ2a”) (25) expressed in rat and rabbit brain (26), but not in rabbit heart (26), have only escalated this requirement, because it has never been shown that the mouse orthologue of this variant is endogenously expressed in the mouse heart.So far, five CaVβ2 variants varying only in the V1 domain have been identified from different species (25, 27, 28) and in human heart these variants have been obtained mainly by RT-PCR approaches (29, 30). In contrast, there is little information on the CaVβ proteins present in mouse heart, their respective splice variants, and expression ratios. We therefore started to study CaVβ expression in the murine heart using Western blots and cDNA cloning and to reveal their functional impact on LTCCs formed by the murine CaV1.2 protein.  相似文献   

15.
The antifungal protein PAF from Penicillium chrysogenum exhibits growth-inhibitory activity against a broad range of filamentous fungi. Evidence from this study suggests that disruption of Ca2+ signaling/homeostasis plays an important role in the mechanistic basis of PAF as a growth inhibitor. Supplementation of the growth medium with high Ca2+ concentrations counteracted PAF toxicity toward PAF-sensitive molds. By using a transgenic Neurospora crassa strain expressing codon-optimized aequorin, PAF was found to cause a significant increase in the resting level of cytosolic free Ca2+ ([Ca2+]c). The Ca2+ signatures in response to stimulation by mechanical perturbation or hypo-osmotic shock were significantly changed in the presence of PAF. BAPTA [bis-(aminophenoxy)-ethane-N,N,N′,N′-tetraacetic acid], a Ca2+ selective chelator, ameliorated the PAF toxicity in growth inhibition assays and counteracted PAF induced perturbation of Ca2+ homeostasis. These results indicate that extracellular Ca2+ was the major source of these PAF-induced effects. The L-type Ca2+ channel blocker diltiazem disrupted Ca2+ homeostasis in a similar manner to PAF. Diltiazem in combination with PAF acted additively in enhancing growth inhibition and accentuating the change in Ca2+ signatures in response to external stimuli. Notably, both PAF and diltiazem increased the [Ca2+]c resting level. However, experiments with an aequorin-expressing Δcch-1 deletion strain of N. crassa indicated that the L-type Ca2+ channel CCH-1 was not responsible for the observed PAF-induced elevation of the [Ca2+]c resting level. This study is the first demonstration of the perturbation of fungal Ca2+ homeostasis by an antifungal protein from a filamentous ascomycete and provides important new insights into the mode of action of PAF.The secreted antifungal protein PAF from Penicillium chrysogenum is a small-molecular-mass (6.2 kDa), cationic, and cysteine-rich peptide that inhibits the growth of numerous filamentous fungi (1416, 21). It belongs to a family of antifungal peptides which show—despite considerable amino acid homology—significant differences in species specificity and modes of action (reviewed in reference 27). Importantly, the primary structures of these antifungals show no similarity to those from higher eukaryotes, e.g., plants, insects, or mammals (see reference 28 for a detailed review on parallels with and differences between PAF and antimicrobial proteins from higher eukaryotes and their mechanisms of action). Apart from the Aspergillus giganteus-derived antifungal protein AFP (19, 45; reviewed in reference 29), PAF is one of the best-studied peptides of this protein family. We have shown that PAF causes rapid hyperpolarization of the plasma membrane at hyphal tips, increased K+ efflux, induction of oxidative stress, and apoptotic cell death (21, 25) and that PAF is internalized by hyphae of PAF-sensitive fungi (33). Furthermore, we have shown that PAF interferes with at least two signaling cascades, the protein kinase C/mitogen-activated protein (MAP) kinase and the cyclic AMP (cAMP)/protein kinase A pathways, which play a role in mediating PAF toxicity (5). However, it still has to be elucidated in more detail how these PAF-dependent effects are linked.Recent evidence indicated that the ionic strength of the growth medium interferes with the antifungal activity of PAF (21). It is possible that Ca2+ ions may play a major role in influencing protein toxicity in an analogous way to the Ca2+-dependent mode of action of antifungal plant defensins. Supplementation of the test medium with low concentrations of CaCl2 (1 to 5 mM) reversed the antifungal activity of plant defensins (34, 4344). The defensins RsAFP2 from the seeds of Raphanus sativus and DmAMP1 from the seeds of Dahlia merckii induced K+ efflux and Ca2+ uptake in Neurospora crassa and caused alkalinization of the growth medium (46). Another seed defensin, MsDef1 from Medicago sativa, was reported to cause Ca2+ influx and the inhibition of mammalian L-type Ca2+ channels, similar to the Ustilago maydis killer toxin KP4 (13, 40).Ca2+ is a universal intracellular second messenger in eukaryotic cells (4, 36). In fungi, there is evidence for Ca2+ signaling regulating numerous processes, including spore germination, tip growth, hyphal branching, sporulation, infection structure differentiation, and circadian clocks, as well as responses to osmotic stress, heat shock, mechanical stimuli, oxidative stress, and electrical fields (7, 11, 17, 20, 22, 3031, 3839, 41, 49). Ca2+ signaling typically involves transient increases in intracellular Ca2+ concentrations originating from the extracellular medium and/or mobilization from internal stores (4, 23, 3637). Little was known about Ca2+ dynamics in filamentous fungi until a new method based on the Ca2+-sensitive photoprotein aequorin was adapted, allowing routine and easy measurement of intracellular calcium dynamics in these organisms (3, 30).The aim of our study was to investigate the effect of PAF on the level of cytosolic free Ca2+ ([Ca2+]c) in the PAF-sensitive target organism N. crassa to characterize its effects on stimulus-specific Ca2+ signatures and to define the Ca2+ sources responsible for the perturbation of Ca2+ homeostasis. By obtaining [Ca2+]c measurements in living cells expressing the bioluminescent Ca2+ reporter aequorin, we provide novel insights into the mode of action of this biotechnologically promising antifungal protein.  相似文献   

16.
All vertebrate cells regulate their cell volume by activating chloride channels of unknown molecular identity, thereby activating regulatory volume decrease. We show that the Ca2+-activated Cl channel TMEM16A together with other TMEM16 proteins are activated by cell swelling through an autocrine mechanism that involves ATP release and binding to purinergic P2Y2 receptors. TMEM16A channels are activated by ATP through an increase in intracellular Ca2+ and a Ca2+-independent mechanism engaging extracellular-regulated protein kinases (ERK1/2). The ability of epithelial cells to activate a Cl conductance upon cell swelling, and to decrease their cell volume (regulatory volume decrease) was dependent on TMEM16 proteins. Activation of ICl,swell was reduced in the colonic epithelium and in salivary acinar cells from mice lacking expression of TMEM16A. Thus TMEM16 proteins appear to be a crucial component of epithelial volume-regulated Cl channels and may also have a function during proliferation and apoptotic cell death.Regulation of cell volume is fundamental to all cells, particularly during cell growth and division. External hypotonicity leads to cell swelling and subsequent activation of volume-regulated chloride and potassium channels, to release intracellular ions and to re-shrink the cells, a process termed regulatory volume decrease (RVD)3 (1). Volume-regulated chloride currents (ICl,swell) have dual functions during cell proliferation as well as apoptotic volume decrease (AVD), preceding apoptotic cell death (2). Although ICl,swell is activated in swollen cells to induce RVD, AVD takes place under normotonic conditions to shrink cells (3, 4). Early work suggested intracellular Ca2+ as an important mediator for activation of ICl,swell and volume-regulated K+ channels (5), whereas subsequent studies only found a permissive role of Ca2+ for activation of ICl,swell (6), reviewed in Ref. 1. In addition, a plethora of factors and signaling pathways have been implicated in activation of ICl,swell, making cell volume regulation an extremely complex process (reviewed in Refs. 1, 3, and 7). These factors include intracellular ATP, the cytoskeleton, phospholipase A2-dependent pathways, and protein kinases such as extracellular-regulated kinase ERK1/2 (reviewed in Refs. 1 and 7). Previous approaches in identifying swelling-activated Cl channels have been unsuccessful or have produced controversial data. Thus none of the previous candidates such as pICln, the multidrug resistance protein, or ClC-3 are generally accepted to operate as volume-regulated Cl channels (reviewed in Refs. 8 and 9). Notably, the cystic fibrosis transmembrane conductance regulator (CFTR) had been shown in earlier studies to influence ICl,swell and volume regulation (1012). The variable properties of ICl,swell suggest that several gene products may affect ICl,swell in different cell types.The TMEM16 transmembrane protein family consists of 10 different proteins with numerous splice variants that contain 8–9 transmembrane domains and have predicted intracellular N- and C-terminal tails (13, 1618). TMEM16A (also called ANO1) is required for normal development of the murine trachea (14) and is associated with different types of tumors, dysplasia, and nonsyndromic hearing impairment (13, 15). TMEM16A has been identified as a subunit of Ca2+-activated Cl channels that are expressed in epithelial and non-epithelial tissues (1618). Interestingly, members of the TMEM16 family have been suggested to play a role in osmotolerance in Saccharomyces cerevisiae (19). Here we show that TMEM16 proteins also contribute to ICl,swell and regulatory volume decrease.  相似文献   

17.
Elevated intracellular Ca2+ ([Ca2+]i) inhibition of NHE3 is reconstituted by NHERF2, but not NHERF1, by a mechanism involving the formation of multiprotein signaling complexes. To further evaluate the specificity of the NHERF family in calcium regulation of NHE3 activity, the current study determined whether NHERF3 reconstitutes elevated [Ca2+]i regulation of NHE3. In vitro, NHERF3 bound the NHE3 C terminus between amino acids 588 and 667. In vivo, NHE3 and NHERF3 associate under basal conditions as indicated by co-immunoprecipitation, confocal microscopy, and fluorescence resonance energy transfer. Treatment of PS120/NHE3/NHERF3 cells, but not PS120/NHE3 cells, with the Ca2+ ionophore, 4-bromo-A23187 (0.5 μm): 1) inhibited NHE3 Vmax activity; 2) decreased NHE3 surface amount; 3) dissociated NHE3 and NHERF3 at the plasma membrane by confocal immunofluorescence and fluorescence resonance energy transfer. Similarly, in Caco-2BBe cells, NHERF3 and NHE3 colocalized in the BB under basal conditions but after elevation of [Ca2+]i by carbachol, this overlap was abolished. NHERF3 short hairpin RNA knockdown (>50%) in Caco-2BBe cells significantly reduced basal NHE3 activity by decreasing BB NHE3 amount. Also, carbachol-mediated inhibition of NHE3 activity was abolished in Caco-2BBe cells in which NHERF3 protein expression was significantly reduced. In summary: 1) NHERF3 colocalizes and directly binds NHE3 at the plasma membrane under basal conditions; 2) NHERF3 reconstitutes [Ca2+]i inhibition of NHE3 activity and dissociates from NHE3 in fibroblasts and polarized intestinal epithelial cells with elevated [Ca2+]i; 3) NHERF3 short hairpin RNA significantly reduced NHE3 basal activity and brush border expression in Caco-2BBe cells. These results demonstrate that NHERF3 reconstitutes calcium inhibition of NHE3 activity by anchoring NHE3 basally and releasing it with elevated Ca2+.In normal digestive physiology, the brush border (BB)2 Na+/H+ exchanger, NHE3, mediates the majority of the NaCl and NaHCO3 absorption in the ileum (1). Sequential inhibition and stimulation of NHE3 occur as part of digestive physiology. Short-term regulation of NHE3 activity is achieved through a variety of factors that affect NHE3 turnover number and/or surface expression and often involve a role for the cytoskeleton and accessory proteins, including the multi-PDZ domain containing proteins, NHERF1 and NHERF2 (1, 2). However, many details of this regulation are not understood.The NHERF (Na+/H+ exchanger regulatory factor) family of multi-PDZ domain containing proteins consists of four evolutionarily related members, all of which are expressed in epithelial cells of the mammalian small intestine (2). NHERF1 and NHERF2 have been previously shown to contribute to acute NHE3 stimulation and inhibition (310). Recently, two additional PDZ domain containing proteins, termed NHERF3/PDZK1 and NHERF4/PDZK2/IKEPP, have been demonstrated to possess sequence homology with NHERF1 and NHERF2 (1114). However, unlike NHERF1 and NHERF2, which are comprised of two tandem PDZ domains flanked by a C-terminal ezrin/radixin/moesin binding domain, NHERF3 and NHERF4 consist of four PDZ domains but no other protein-protein interacting domains (12).NHERF3 was initially identified by a yeast two-hybrid screen from a human kidney cDNA library using the membrane-associated protein MAP17, as bait (12). NHERF3 is expressed in the brush border of epithelial cells of the kidney proximal tubule and the small intestine (12). NHERF3 associates with and, in a few cases, has been shown to regulate the activity of multiple apical membrane ion transporters including the cystic fibrosis transmembrane regulator (CFTR), urate anion exchanger 1 (URAT1), sodium-phosphate cotransporter type IIa (NaPiIIa), proton-coupled peptide transporter (PEPT2), and organic cation/carnitine cotransporter (OCTN2) (1519). Furthermore, NHERF3 directly binds the C terminus of NHE3 (20). Recent studies have begun evaluating the effect of NHERF3 on mouse intestinal Na+ and Cl transport. Basal electroneutral sodium absorption was decreased by >40% in the NHERF3 null mouse jejunum (21) and by >80% in the colon (22). In addition, Cinar et al. (22) demonstrated that cAMP and [Ca2+]i inhibition of NHE3 activity was abolished in the NHERF3 null mouse colon. However, the mechanism by which NHERF3 regulates NHE3 activity was not resolved.Several physiological and pathophysiological agonists, acting through [Ca2+]i-induced second messenger systems, are known to inhibit electroneutral NaCl absorption in the small intestine (1, 23). Elevation of [Ca2+]i has previously been demonstrated to inhibit NHE3 activity in a NHERF2-, but not NHERF1-dependent manner (5). NHERF2 regulation of NHE3 involves the formation of multiprotein complexes at the plasma membrane that include NHE3, NHERF2, α-actinin-4, and PKCα, which induce endocytic removal of NHE3 from the plasma membrane by a PKC-dependent mechanism (5, 24). Because multiple PDZ proteins exist in the apical pole of epithelial cells (2), the current study was designed to determine whether NHERF3 could reconstitute Ca2+ regulation of NHE3 activity and to define how that occurred.  相似文献   

18.
Intracellular signaling systems of G protein-coupled receptors are well established, but their role in paracrine regulation of adjacent cells is generally considered as a tissue-specific mechanism. We have shown previously that AT1 receptor (AT1R) stimulation leads to diacylglycerol lipase-mediated transactivation of co-expressed CB1Rs in Chinese hamster ovary cells. In the present study we detected a paracrine effect of the endocannabinoid release from Chinese hamster ovary, COS7, and HEK293 cells during the stimulation of AT1 angiotensin receptors by determining CB1 cannabinoid receptor activity with bioluminescence resonance energy transfer-based sensors of G protein activation expressed in separate cells. The angiotensin II-induced, paracrine activation of CB1 receptors was visualized by detecting translocation of green fluorescent protein-tagged β-arrestin2. Mass spectrometry analyses have demonstrated angiotensin II-induced stimulation of 2-arachidonoylglycerol production, whereas no increase of anandamide levels was observed. Stimulation of Gq/11-coupled M1, M3, M5 muscarinic, V1 vasopressin, α1a adrenergic, B2 bradykinin receptors, but not Gi/o-coupled M2 and M4 muscarinic receptors, also led to paracrine transactivation of CB1 receptors. These data suggest that, in addition to their retrograde neurotransmitter role, endocannabinoids have much broader paracrine mediator functions during activation of Gq/11-coupled receptors.Hormones, neurotransmitters, and other chemical mediators acting on G protein-coupled receptors (GPCRs)2 exert their effects on the target cells by stimulating G protein-dependent and independent intracellular signaling pathways (14). Activation of Gq/11 protein-coupled receptors causes phospholipase C activation, which produces inositol-trisphosphate and diacylglycerol from phosphatidylinositol (4,5)-bisphosphate, leading to Ca2+-signal generation and protein kinase C activation. However, the concerted response of tissues to chemical mediators frequently also involves the activation of cells adjacent to the target cells, due to the release of paracrine mediators. A well known example is NO, which can be released from activated endothelial cells to cause relaxation of adjacent vascular smooth muscle cells. Lipid mediators can also act as intercellular messengers. For example, endocannabinoids released from postsynaptic neurons after depolarization act as retrograde transmitters by binding to and stimulating presynaptic cannabinoid receptors, which leads to inhibition of γ-aminobutyric acid release (an event termed depolarization-induced suppression of inhibition, DSI) (57).Cannabinoid receptors were first identified based on their ability to selectively recognize marijuana analogs. To date, two cannabinoid receptors have been identified by molecular cloning, CB1 and CB2 receptors (CB1R and CB2R, respectively) (5, 8, 9), although additional GPCRs have also been proposed to function as cannabinoid receptors (10, 11). Cannabinoid receptors also recognize certain lipids present in animal tissues termed endocannabinoids, such as arachidonylethanolamide (anandamide), 2-arachidonoylglycerol (2-AG), and 2-arachidonoylglyceryl ether (noladin ether) (7, 1216). In adult and fetal neural tissues, the two major endocannabinoids, anandamide and 2-AG, are produced on demand, usually after depolarization of postsynaptic cells or following stimulation of Gq-coupled metabotropic glutamate or muscarinic acetylcholine receptors (7, 12, 1720). Enzymes responsible for 2-AG production and metabolism in tissues are localized to well defined structures at synapses, near the axon terminals of CB1R-expressing cells (5, 7). In contrast, in peripheral tissues baseline levels of endocannabinoid production usually manifest as “endocannabinoid tone,” with poorly understood localization of the various components of the endocannabinoid system. 2-AG levels in brain homogenates and in many peripheral tissues are near its Kd for the CB1R (19), suggesting that function of endocannabinoids may not be limited to localized synaptic signaling.There is mounting evidence that endocannabinoids play important roles in peripheral cardiovascular, inflammatory, intestinal, and metabolic regulation (2124). 2-AG is produced by diacylglycerol-lipase (DAGL) after cleavage of the fatty-acid in the sn-1 position of diacylglycerol (DAG) (19, 25). Phospholipase C activation by Gq/11 protein-coupled receptors produces DAG, which can serve as a substrate for DAGL. Plasma membrane phosphoinositides are enriched in arachidonic acid in the sn-2 position (26), and DAGL is expressed ubiquitously (27), which suggests that phospholipase C-mediated cleavage of polyphosphoinositides may routinely lead to the formation of 2-AG. In accordance with this hypothesis, we have recently shown that angiotensin II- (Ang II)-mediated activation of the Gq/11-coupled AT1 angiotensin receptor (AT1R) leads to DAGL-dependent activation of CB1Rs expressed in Chinese hamster ovary (CHO) cells (28).Here our aim has been to examine the possibility that 2-AG serves as a common paracrine signal generated via activation of Gq/11 protein-coupled, Ca2+-mobilizing receptors. Accordingly, we co-expressed CB1Rs and BRET-based sensors of G protein activation in CHO cells, and used these cells to detect endocannabinoid release from adjacent cells that express AT1R or other Ca2+-mobilizing GPCRs. We have further shown that activation of AT1R by Ang II increases 2-AG levels in CHO cells. These findings suggest that 2-AG is commonly released following activation of Ca2+-mobilizing GPCRs and serves as a paracrine signal to activate CB1R in neighboring cells.  相似文献   

19.
20.
Recent studies identified two main components of store-operated calcium entry (SOCE): the endoplasmic reticulum-localized Ca2+ sensor protein, STIM1, and the plasma membrane (PM)-localized Ca2+ channel, Orai1/CRACM1. In the present study, we investigated the phosphoinositide dependence of Orai1 channel activation in the PM and of STIM1 movements from the tubular to PM-adjacent endoplasmic reticulum regions during Ca2+ store depletion. Phosphatidylinositol 4,5-bisphosphate (PtdIns(4,5)P2) levels were changed either with agonist stimulation or by chemically induced recruitment of a phosphoinositide 5-phosphatase domain to the PM, whereas PtdIns4P levels were decreased by inhibition or down-regulation of phosphatidylinositol 4-kinases (PI4Ks). Agonist-induced phospholipase C activation and PI4K inhibition, but not isolated PtdIns(4,5)P2 depletion, substantially reduced endogenous or STIM1/Orai1-mediated SOCE without preventing STIM1 movements toward the PM upon Ca2+ store depletion. Patch clamp analysis of cells overexpressing STIM1 and Orai1 proteins confirmed that phospholipase C activation or PI4K inhibition greatly reduced ICRAC currents. These results suggest an inositide requirement of Orai1 activation but not STIM1 movements and indicate that PtdIns4P rather than PtdIns(4,5)P2 is a likely determinant of Orai1 channel activity.Store-operated Ca2+ entry (SOCE)3 is a ubiquitous Ca2+ entry pathway that is regulated by the Ca2+ content of the endoplasmic reticulum (ER) (1). SOCE has been identified as the major route of Ca2+ entry during activation of cells of the immune system such as T cells and mast cells (2, 3), and it is also present and functionally important in other cells such as platelets (4) and developing myotubes (5). The long awaited mechanism of how the ER luminal Ca2+ content is sensed and the information transferred to the plasma membrane (PM) has been clarified recently after identification of the ER Ca2+ sensor proteins STIM1 and -2 (6, 7) and the PM Ca2+ channels Orai1, -2, and -3 (810). According to current views, a decrease in the ER Ca2+ concentration is sensed by the luminal EF-hand of the single-transmembrane STIM proteins causing their multimerization. This oligomerization occurs in the tubular ER, where it promotes the interaction of the cytoplasmic C termini of STIM with PM components and association with the PM-localized Orai channels, causing both their clustering and activation in the PM (reviewed recently in Refs. 1113). Analysis of the interacting domains within the STIM1 and Orai1 proteins suggests that the cytoplasmic domain of STIM1 is necessary and sufficient to activate Orai1 (14), whereas the latter requires its C-terminal membrane-adjacent cytoplasmic tail to be fully activated by the STIM proteins (15, 16). Both STIM1 and -2 contain a polybasic segment in their C termini, and such regions are often responsible for the PM localization of proteins (mostly of the small GTP-binding protein class) via interaction with anionic phospholipids such as phosphatidylserine or PtdIns(4,5)P2 (17). However, the role of this domain in STIM1 function(s) remains controversial. Deletion of the polybasic tail is reported to prevent PM association but not clustering of STIM1 upon ER store depletion (18). In other studies, truncated STIM1 lacking the polybasic domain shows only slightly altered activation (15) or inactivation (19) kinetics without major defects in supporting Orai1-mediated Ca2+ influx. The most recent studies identify the minimal Orai1 activation domain in STIM1 (20, 21) and find that the polybasic domain is not essential for this function but makes electrostatic interaction with classical transient receptor potential channels (22).PM phosphoinositides have been widely reported as regulators of the activity of several ion channels and transporters (23). However, only a few studies have addressed the inositide requirement of SOCE and none specifically that of the Orai1-mediated Ca2+ entry process. Sensitivity of SOCE to phosphatidylinositol 3-kinases (PI3K) inhibitors has been reported, but this required concentrations that suggested inhibition of targets other than PI3Ks, possibly myosin light chain kinase or the type-III PI4Ks (4, 2426). Here we have described studies addressing the role of PM phosphoinositides in STIM1 movements as well as in Orai1 channel gating. Our results show that phosphoinositides do not have a major role in the prominent reorganization of STIM1 after Ca2+ store depletion but suggest a function of PtdIns4P rather than PtdIns(4,5)P2 in supporting the Orai1-mediated Ca2+ entry process.  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