首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
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   

2.
The TRPV4 (transient receptor potential vanilloid 4) ion channel, a member of the vanilloid subfamily of the transient receptor potential channels, is activated by membrane stretch, by non-noxious warm temperatures, and by a range of chemical activators. In the present study we examined the role of phosphorylation in modulating the activation of TRPV4. We expressed TRPV4 in HEK293 cells and activated the channel by cell swelling in a hypotonic solution. TRPV4 channel activation and serine phosphorylation were enhanced by exposure to the protein kinase C (PKC) activator phorbol 12-myristate 13-acetate or by application of bradykinin, which activates PKC via a G-protein-coupled mechanism. The enhancement was inhibited by the PKC inhibitors staurosporine, bisindolylmaleimide I, and rottlerin or by mutation of the serine/threonine residues Ser162, Thr175, and Ser189. The adenylate cyclase activator forskolin also enhanced activation of TRPV4, and the enhancement was antagonized by the selective cyclic AMP-dependent protein kinase (PKA) inhibitor H89 or by mutation of serine residue Ser824. Sensitization of TRPV4 by both PKC and PKA depended on the scaffolding protein AKAP79, because channel activation and phosphorylation were enhanced by co-transfection of AKAP79 and were antagonized by removal of AKAP79 using small interfering RNA. We conclude that the serine/threonine kinases PKC and PKA enhance activation of the TRPV4 ion channel by phosphorylation at specific sites and that phosphorylation depends on assembly of PKC and PKA by AKAP79 into a signaling complex with TRPV4.TRPV4 was cloned from kidney, hypothalamus, and auditory epithelium and was given a number of names: OTRPC4 (Osm-9-like TRP channel 4) (1), VR-OAC (2), TRP12 (3), and VRL-2 (vanilloid receptor-like channel 2) (4). The gene for human TRPV4 is located on chromosome 12q23-q24.1 and has 15 exons, which code for a full-length protein with 871 amino acids. TRPV4 is a member of the transient receptor potential vanilloid subfamily of TRP2 channels, and like other members of this subfamily, it is a polymodal receptor activated by a wide variety of stimuli. TRPV4 is strongly expressed in kidney and is activated by hypotonicity, which has led to the suggestion that TRPV4 is an osmosensor important in regulating body fluid levels (2, 59). However, TRPV4 is also activated by innocuous heat with a threshold of >27 °C (6, 10, 11), by the phorbol ester 4α-phorbol 12,13-didecanoate (12, 13), by low pH (14), by endocannabinoids and arachidonic acid metabolites (15, 16), by the active compound, bisandrographolide A, of Andrographis paniculata, a Chinese herbal plant (17), and by nitric oxide (18). TRPV4 is expressed in a broad range of tissues, including lung, spleen, kidney, testis, fat, brain, cochlea, skin, smooth muscle, liver, and vascular endothelium (13); in the lamina terminalis of the mouse brain; in neurons of the arched vascular organ of the lamina terminalis; and in the median preoptic area, the optic chiasm, neurons of the subfornical organ, the ventral hippocampal commissure, anterior hypothalamic structures, and ependymal cells of the choroid plexus in the lateral ventricles, and dorsal root ganglia neurons (13). The broad spectrum of activators and the wide distribution of TRPV4 suggest that the functions of TRPV4 extend beyond osmosensation.TRPV4 has been proposed to play a role in the mechanical hyperalgesia that is generated by the concerted action of inflammatory mediators present in inflamed tissues (19). After tissue injury, inflammatory mediators such as bradykinin, prostaglandin E2, 5-hydroxytryptamine, and histamine directly sensitize primary afferent neurons, resulting in hyperalgesia (reviewed in Ref. 20). Important intracellular signaling molecules contributing to inflammatory hyperalgesia include protein kinase C (PKC) (21, 22) and cyclic AMP-dependent protein kinase (PKA) (23). For example, the activation of the Gq-coupled B1 and B2 receptors by bradykinin leads to the release of a range of potential intracellular messengers, with a substantial body of evidence favoring the idea that the temperature threshold of TRPV1 is lowered by PKCϵ-mediated phosphorylation (21, 22, 24, 25). PKA, like PKC, is a critical intracellular signaling molecule mediating inflammatory hyperalgesia (26). In sensory neurons prostaglandin E2 activates both the EP1 receptor, which is Gq-coupled and therefore activates PKC, and the EP4 receptor, which is Gs-coupled and therefore activates PKA. Cyclic AMP analogues, the adenylate cyclase activator forskolin (FSK) or phosphodiesterase inhibitors enhance the mechanical and thermal hyperalgesic effects of prostaglandin E2 (2729). Thus PKC and PKA have vital roles to play in the process of inflammatory hyperalgesia.The speed and specificity of the action of kinases is in many cases enhanced by binding to scaffolding proteins, which preassemble the kinases into signaling complexes with their target substrates. The AKAP (a kinase-anchoring protein) family of scaffolding proteins was originally named for their ability to target PKA to appropriate substrates but are now known to assemble a wide range of kinases and phosphatases into signaling complexes with appropriate targets (30). A number of ion channels are subject to modulation by AKAPs, including glutamate receptors, calcium channels, and the M-type potassium channels (3134). The heat-activated ion channel TRPV1, a member of the same subfamily as TRPV4, has recently been shown to be assembled into a signaling complex with PKA, PKC, and PP2B by AKAP79, and the sensitization of TRPV1 by PKC and PKA is critically reliant on binding to AKAP79 (35). The present study shows that PKC and PKA activation can sensitize TRPV4 to mechanical stimuli, identifies the relevant phosphorylation sites, and shows that the scaffolding protein AKAP79 plays a critical role in sensitization of TRPV4.  相似文献   

3.
Vascular smooth muscle Kv1 delayed rectifier K+ channels (KDR) containing Kv1.2 control membrane potential and thereby regulate contractility. Vasodilatory agonists acting via protein kinase A (PKA) enhance vascule smooth muscle Kv1 activity, but the molecular basis of this regulation is uncertain. We characterized the role of a C-terminal phosphorylation site, Ser-449, in Kv1.2 expressed in HEK 293 cells by biochemical and electrophysiological methods. We found that 1) in vitro phosphorylation of Kv1.2 occurred exclusively at serine residues, 2) one major phosphopeptide that co-migrated with 449pSASTISK was generated by proteolysis of in vitro phosphorylated Kv1.2, 3) the peptide 445KKSRSASTISK exhibited stoichiometric phosphorylation by PKA in vitro, 4) matrix-assisted laser desorption ionization time-of-flight (MALDI-TOF) mass spectroscopy (MS) and MS/MS confirmed in vitro Ser-449 phosphorylation by PKA, 5) in situ phosphorylation at Ser-449 was detected in HEK 293 cells by MALDI-TOF MS followed by MS/MS. MIDAS (multiple reaction monitoring-initiated detection and sequencing) analysis revealed additional phosphorylated residues, Ser-440 and Ser-441, 6) in vitro 32P incorporation was significantly reduced in Kv1.2-S449A, Kv1.2-S449D, and Kv1.2-S440A/S441A/S449A mutant channels, but Kv1.2-S440A/S441A was identical to wild-type Kv1.2 (Kv1.2-WT), and 7) bath applied 8-Br-cAMP or dialysis with PKA catalytic subunit (cPKA) increased Kv1.2-WT but not Kv1.2-S449A current amplitude. cPKA increased Kv1.2-WT current in inside-out patches. Rp-CPT-cAMPS reduced Kv1.2-WT current, blocked the increase due to 8-Br-cAMP, but had no effect on Kv1.2-S449A. cPKA increased current due to double mutant Kv1.2-S440A/S441A but had no effect on Kv1.2-S449D or Kv1.2-S440A/S441A/S449A. We conclude that Ser-449 in Kv1.2 is a site of PKA phosphorylation and a potential molecular mechanism for Kv1-containing KDR channel modulation by agonists via PKA activation.Voltage-gated K+ channels (Kv)4 composed of members of the Kv1 family (Kv1) are expressed in vascular smooth muscle (VSM) and other excitable cells, where they play an important role in controlling membrane potential, a key mechanism for regulation of contraction and arterial diameter (1). Steady-state Ca2+ influx through voltage-gated Ca2+ channels, resulting in elevated cytosolic Ca2+ concentration and contraction, requires sustained depolarization. In contrast, Kv1 channel activation induces hyperpolarization, reduces voltage-gated Ca2+ channel open probability, and promotes relaxation (13).Phosphorylation of ion channels by protein kinases is an important mechanism by which ion channel gating (opening and closing) and, thereby, membrane potential are modulated by cellular signaling pathways. VSM Kv1 channels containing Kv1.2-Kv1.5 (46) alone or with Kv1.6 (7) contribute to control of arterial diameter by vasoconstrictor and vasodilator stimuli (813). For example, hyperpolarization and relaxation induced by elevations in intracellular (cAMP) are suppressed by the block of Kv1 channels (1418), and the activity of Kv1-containing VSM KDR is enhanced by protein kinase A (PKA) (1013, 1922). However, the molecular basis of this regulation is unclear.Previous studies showed that gating of Kv1.2 channels expressed in Xenopus oocytes is stimulated by PKA (23), similar to VSM KDR channels, which contain this pore-forming Kv1 subunit (713). There are six potential PKA phosphorylation sites within the cytoplasmic regions of Kv1.2 (see Fig. 1, A and B) based on the consensus motif Arg/Lys-Xaa-Xaa-Ser/Thr (24). An N-terminal residue, Thr-46, was proposed to be the site of PKA phosphorylation, as mutation to valine (T46V) prevented modulation by β-adrenoreceptor activation or exposure to the catalytic subunit of PKA (cPKA) (23). However, this conclusion has been questioned based on structural analysis of the N-terminal domain and the positive shift in voltage-dependent gating produced by the T46V mutation (25). Sites Ser-440, Ser-441, and Ser-449 near the C terminus were recently identified to be phosphorylated in Kv1.2 immunoprecipitates derived from brain as well as transfected human embryonic kidney (HEK 293) and COS-1 cells using a proteomic approach (26). Phosphorylation of these residues was suggested to be important for trafficking to the cell membrane (26), but the kinase(s) involved was not identified. cAMP/PKA-dependent signaling was reported to alter surface expression of Kv1.2, but phosphorylation at Ser-440 and Ser-449 was not required (27).Open in a separate windowFIGURE 1.Identification of potential PKA consensus phosphorylation sites in Kv1.2 and immunoprecipitation of Myc-tagged wild-type and mutant Kv1.2. A, schematic of Kv1.2 with candidate PKA phosphorylation sites indicated by asterisks, and Ser/Thr residues relevant to the study are labeled. B, amino acid sequence of Kv1.2 indicating transmembrane domains (boxes), cytoplasmic domains (italics) assessed by MIDAS for phosphorylated residues, candidate PKA consensus sites (bold), and potential phosphorylated Ser/Thr residues (underlined). C, immunoblot (IB) analysis of untagged Kv1.2-WT and Myc-tagged WT and mutant constructs using anti-Kv1.2. D, immunoblot identification of untagged Kv1.2-WT and Myc-tagged WT and mutant constructs by anti-Kv1.2 in anti-Myc immunoprecipitates (IP).In the present study we used a combination of biochemical, mutagenesis, and electrophysiological analyses to 1) determine whether Kv1.2 is a substrate for PKA, 2) identify the PKA phosphorylation site(s) involved, and 3) determine whether phosphorylation of the identified residue(s) mediates PKA-dependent changes in Kv1.2 current amplitude.  相似文献   

4.
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   

5.
In the past decade, mass-spectrometry-based methods have emerged for the quantitative profiling of dynamic changes in protein phosphorylation, allowing the behavior of thousands of phosphorylation sites to be monitored in a single experiment. However, when one is interested in specific signaling pathways, such shotgun methodologies are not ideal because they lack selectivity and are not cost and time efficient with respect to instrument and data analysis time.Here we evaluate and explore a peptide-centric antibody generated to selectively enrich peptides containing the cAMP-dependent protein kinase (PKA) consensus motif. This targeted phosphoproteomic strategy is used to profile temporal quantitative changes of potential PKA substrates in Jurkat T lymphocytes upon prostaglandin E2 (PGE2) stimulation, which increases intracellular cAMP, activating PKA. Our method combines ultra-high-specificity motif-based immunoaffinity purification with cost-efficient stable isotope dimethyl labeling. We identified 655 phosphopeptides, of which 642 (i.e. 98%) contained the consensus motif [R/K][R/K/X]X[pS/pT]. When our data were compared with a large-scale Jurkat T-lymphocyte phosphoproteomics dataset containing more than 10,500 phosphosites, a minimal overlap of 0.2% was observed. This stresses the need for such targeted analyses when the interest is in a particular kinase.Our data provide a resource of likely substrates of PKA, and potentially some substrates of closely related kinases. Network analysis revealed that about half of the observed substrates have been implicated in cAMP-induced signaling. Still, the other half of the here-identified substrates have been less well characterized, representing a valuable resource for future research.The identification and quantification of protein phosphorylation under system perturbations is an integral part of systems biology (1, 2). The combination of phosphopeptide enrichment (36), stable isotope labeling, and high-resolution mass spectrometry (MS) methods (79) has become the method of choice for the identification of novel phosphorylation sites and for the quantitation of temporal dynamics within signaling networks (10, 11), allowing the behavior of thousands of phosphorylation sites to be studied in a single experiment (10, 12, 13). Nowadays, one of the most commonly adopted high-throughput phosphoproteomics strategies utilizes two consecutive separation steps: (i) an initial fractionation to reduce the sample complexity, and (ii) a phosphopeptide-specific affinity purification. Such techniques include strong cation exchange fractionation under acidic conditions (3), followed by a chelation-based method with the use of metal ions (i.e. immobilized metal ion affinity chromatography (4), metal oxide affinity chromatography (10, 14), or Ti4+ immobilized metal ion affinity chromatography (6)). Alternatives to strong cation exchange for the first sample fractionation step have also been reported, including the use of electrostatic repulsion liquid chromatography (15, 16), which is well suited for the identification of multiply phosphorylated peptides, or hydrophilic interaction chromatography (17).Although the number of detected phosphorylated peptides is nowadays impressive, these kinds of methodologies are still inclined to identify/quantify the more abundant phosphoproteins present in a sample. For example, phosphotyrosine peptides are underrepresented because of their relatively lower abundance.In order to analyze key signaling events that may occur on less abundant phosphoproteins, more targeted approaches, focused on a specific pathway or a specific post-translational modification, are thus still essential. Studies examining post-translational modifications are often based on immunoaffinity purification at the protein or peptide level using dedicated antibodies. Recent examples include the selective enrichment of acetylated lysines (18) and phosphorylated tyrosines (19, 20). More recently, the first specific methods targeting serine/threonine phosphorylation motifs using immune-affinity assays have emerged (21, 22). The advantages of targeted approaches are their potentially higher sensitivity and more specific throughput with, as a consequence, relatively faster and easier data interpretation, which make them attractive for many systems biology applications.Immunoaffinity enrichment can be applied at both the protein and the peptide level, and both have been explored to study protein tyrosine phosphorylation (23). The first one results mainly in information on total protein phosphorylation levels. The detection of the actual phosphoresidue might be hampered by the high content of unmodified peptides derived from the immune-purified phosphoprotein and its binding partners. Immunoprecipitation at the peptide level (20, 24, 25), in contrast, leads to improved phosphosite characterization, with the identification of hundreds of sites, albeit with the loss (generally) of information regarding total protein expression.To profile the dynamic regulation of phosphorylation events via mass spectrometry, stable isotope labeling is often implemented, either with the use of amino acids in cell culture (10) or via chemical peptide labeling of the proteolytic digests (26, 27). To identify low-abundant signaling events, phosphoprotein/phosphopeptide immunoprecipitation is typically performed on several milligrams of material because of the substoichiometric abundance of post-translational modifications. This may hamper the use of expensive isotope-labeling reagents such as iTRAQ or tandem mass tag reagents, given the large amount of chemicals needed. Boersema et al. (28) introduced an alternative sensitive and accurate triplex labeling approach using inexpensive reagents (i.e. formaldehyde) that is much less limited in terms of the sample type or amount. We combined this latter stable-isotope dimethyl labeling approach (2729) with highly specific antibodies raised against a set of cAMP-dependent protein kinase (PKA) phosphorylated substrates as based on the current literature (11, 3034). It is generally accepted that PKA phosphorylates sites with the reasonably stringent consensus motif [R/K][R/K/X]X[pS/pT]. It should be noted that this consensus motif resembles somewhat the motifs of other AGC kinases (e.g. Akt, PKG, PKC).The basicity of the PKA motifs may hamper their analysis via MS-based proteomics, especially when trypsin is used as a protease, as the peptides may become too small to be sequenced. The use of trypsin is also unfavorable in the approach presented here when attempting to immunoprecipitate peptides bearing the PKA motif. Therefore, we decided to use Lys-C in order to keep the (dominant (RRX[pS/pT])) phosphorylated motif intact. To enhance identification, we applied decision-tree MS/MS technology (9), which makes use of HCD and ETD for more efficient fragmentation, higher mass accuracy in tandem MS mode, and less background noise (35).We applied this method to screen the response of Jurkat T cells to prostaglandin E2 (PGE2) treatment. PGE2 is a potent inflammatory mediator that plays an important role in several immune-regulatory actions (36). It is produced by many different cell types, including tumor cells, where carcinogenesis is associated with chronic inflammatory responses (37). PGE2 signaling in T cells is initiated by its binding to the G protein–coupled receptors EP1, -2, -3, and -4. Signaling pathways that are initiated by PGE2 are for the most part under control of the second messenger cyclic adenosine monophosphate (cAMP),1 which is generated from ATP by adenylyl cyclase when PGE2 binds to EP2 or EP4 receptors. One of the primary targets of cAMP is PKA—cAMP binding releases the catalytic subunit activating the kinase. In the current study, we efficiently enriched close to 650 phosphopeptides containing the [R/K][R/K/X]X[pS/pT] consensus motif. Almost all these sites were absent in a recently reported comprehensive phosphoproteomics dataset of Jurkat T cells (12), compiled using shotgun strong cation exchange–immobilized metal ion affinity chromatography analysis and containing ∼10,500 phosphorylation sites, illustrative of the complementarity and selectivity of our approach. The qualitative and quantitative data presented here provide a wide-ranging and credible resource of likely PKA substrates. Network analysis confirmed several established cAMP-dependent signaling nodes in our dataset, although most identified potential PKA substrates are “novel” (i.e. not previously reported and/or linked to PKA). Therefore, the dataset presented here can be considered as a comprehensive and reliable resource for future research into cAMP-related signaling.  相似文献   

6.
The multifunctional autoprocessing repeats-in-toxin (MARTX) toxin of Vibrio cholerae causes destruction of the actin cytoskeleton by covalent cross-linking of actin and inactivation of Rho GTPases. The effector domains responsible for these activities are here shown to be independent proteins released from the large toxin by autoproteolysis catalyzed by an embedded cysteine protease domain (CPD). The CPD is activated upon binding inositol hexakisphosphate (InsP6). In this study, we demonstrated that InsP6 is not simply an allosteric cofactor, but rather binding of InsP6 stabilized the CPD structure, facilitating formation of the enzyme-substrate complex. The 1.95-Å crystal structure of this InsP6-bound unprocessed form of CPD was determined and revealed the scissile bond Leu3428–Ala3429 captured in the catalytic site. Upon processing at this site, CPD was converted to a form with 500-fold reduced affinity for InsP6, but was reactivated for high affinity binding of InsP6 by cooperative binding of both a new substrate and InsP6. Reactivation of CPD allowed cleavage of the MARTX toxin at other sites, specifically at leucine residues between the effector domains. Processed CPD also cleaved other proteins in trans, including the leucine-rich protein YopM, demonstrating that it is a promiscuous leucine-specific protease.Multifunctional-autoprocessing repeats-in-toxin (MARTX)3 toxins are a family of large bacterial protein toxins with conserved repeat regions at the N and C termini that are predicted to transfer effector domains located between the repeats across the eukaryotic cell plasma membrane (1). The best characterized MARTX is the >450-kDa secreted virulence-associated MARTX of Vibrio cholerae. This toxin causes disassembly of the actin cytoskeleton and enhances V. cholerae colonization of the small intestine, possibly by facilitating evasion of phagocytic cells (2, 3). The central region of the V. cholerae MARTX toxin contains four discrete domains: the actin cross-linking domain (ACD) that introduces lysine-glutamate cross-links between actin protomers (4, 5), the Rho-inactivating domain (RID) that disables small Rho GTPases (6), an αβ hydrolase of unknown function (1), and an autoprocessing cysteine protease domain (CPD) (7, 8).The CPD is a 25-kDa domain found in all MARTX toxins located just before the start of the C-terminal repeats (7, 8). This domain is activated for autoproteolysis upon binding inositol hexakisphosphate (InsP6) (7), a molecule ubiquitously present in eukaryotic cell cytosol (911), but absent in extracellular spaces and bacteria. Thus, autocatalytic processing would not occur until after translocation of the CPD and effector domains is completed. In the context of the holotoxin, catalytic residue Cys3568 was found to be essential for the toxin to induce efficient actin cross-linking by the ACD and Rho inactivation by the RID, demonstrating that autoprocessing is essential for MARTX to induce cell rounding (8).While it is clear that InsP6 activates the CPD and that autoprocessing is essential for MARTX function (7), the mechanism by which InsP6 activates CPD is not well understood. Furthermore, only one processing site at Leu3428–Ala3429 has been identified, although multiple processing events would be required to release each effector independently. In fact, after autoprocessing at Leu3428–Ala3429, CPD is reported to adopt a conformation with reduced affinity for InsP6 (7), raising questions as to how the protease might process MARTX at other sites.We present here the structure of the pre-processed form of the V. cholerae MARTX CPD bound to InsP6. Our results demonstrate that autoprocessing is activated by rearrangement of a β-hairpin loop upon InsP6 binding that locks the N terminus of the CPD in the active site, facilitating hydrolysis of the Leu3428–Ala3429 peptide bond. After autoprocessing, CPD adopts a post-processing form that has poor affinity for InsP6 and thus must be cooperatively reactivated for high affinity binding of InsP6 by association of a new substrate. As a consequence, we are able to demonstrate how CPD cleaves MARTX toxin between effector domains and releases them from the large toxin resulting in increased catalytic activity of the effectors.  相似文献   

7.
Collapsin response mediator protein 2 (CRMP2) is an intracellular protein that mediates signaling of Semaphorin3A (Sema3A), a repulsive axon guidance molecule. Fyn, a Src-type tyrosine kinase, is involved in the Sema3A signaling. However, the relationship between CRMP2 and Fyn in this signaling pathway is still unknown. In our research, we demonstrated that Fyn phosphorylated CRMP2 at Tyr32 residues in HEK293T cells. Immunohistochemical analysis using a phospho-specific antibody at Tyr32 of CRMP showed that Tyr32-phosphorylated CRMP was abundant in the nervous system, including dorsal root ganglion neurons, the molecular and Purkinje cell layer of adult cerebellum, and hippocampal fimbria. Overexpression of a nonphosphorylated mutant (Tyr32 to Phe32) of CRMP2 in dorsal root ganglion neurons interfered with Sema3A-induced growth cone collapse response. These results suggest that Fyn-dependent phosphorylation of CRMP2 at Tyr32 is involved in Sema3A signaling.Collapsin response mediator proteins (CRMPs)4 have been identified as intracellular proteins that mediate Semaphorin3A (Sema3A) signaling in the nervous system (1). CRMP2 is one of the five members of the CRMP family. CRMPs also mediate signal transduction of NT3, Ephrin, and Reelin (24). CRMPs interact with several intracellular molecules, including tubulin, Numb, kinesin1, and Sra1 (58). CRMPs are involved in axon guidance, axonal elongation, cell migration, synapse maturation, and the generation of neuronal polarity (1, 2, 4, 5).CRMP family proteins are known to be the major phosphoproteins in the developing brain (1, 9). CRMP2 is phosphorylated by several Ser/Thr kinases, such as Rho kinase, cyclin-dependent kinase 5 (Cdk5), and glycogen synthase kinase 3β (GSK3β) (2, 1013). The phosphorylation sites of CRMP2 by these kinases are clustered in the C terminus and have already been identified. Rho kinase phosphorylates CRMP2 at Thr555 (10). Cdk5 phosphorylates CRMP2 at Ser522, and this phosphorylation is essential for sequential phosphorylations by GSK3β at Ser518, Thr514, and Thr509 (2, 1113). These phosphorylations disrupt the interaction of CRMP2 with tubulin or Numb (2, 3, 13). The sequential phosphorylation of CRMP2 by Cdk5 and GSK3β is an essential step in Sema3A signaling (11, 13). Furthermore, the neurofibrillary tangles in the brains of people with Alzheimer disease contain hyperphosphorylated CRMP2 at Thr509, Ser518, and Ser522 (14, 15).CRMPs are also substrates of several tyrosine kinases. The phosphorylation of CRMP2 by Fes/Fps and Fer has been shown to be involved in Sema3A signaling (16, 17). Phosphorylation of CRMP2 at Tyr479 by a Src family tyrosine kinase Yes regulates CXCL12-induced T lymphocyte migration (18). We reported previously that Fyn is involved in Sema3A signaling (19). Fyn associates with PlexinA2, one of the components of the Sema3A receptor complex. Fyn also activates Cdk5 through the phosphorylation at Tyr15 of Cdk5 (19). In dorsal root ganglion (DRG) neurons from fyn-deficient mice, Sema3A-induced growth cone collapse response is attenuated compared with control mice (19). Furthermore, we recently found that Fyn phosphorylates CRMP1 and that this phosphorylation is involved in Reelin signaling (4). Although it has been shown that CRMP2 is involved in Sema3A signaling (1, 11, 13), the relationship between Fyn and CRMP2 in Sema3A signaling and the tyrosine phosphorylation site(s) of CRMPs remain unknown.Here, we show that Fyn phosphorylates CRMP2 at Tyr32. Using a phospho-specific antibody against Tyr32, we determined that the residue is phosphorylated in vivo. A nonphosphorylated mutant CRMP2Y32F inhibits Sema3A-induced growth cone collapse. These results indicate that tyrosine phosphorylation by Fyn at Tyr32 is involved in Sema3A signaling.  相似文献   

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

9.
Stathmin is an important regulator of microtubule polymerization and dynamics. When unphosphorylated it destabilizes microtubules in two ways, by reducing the microtubule polymer mass through sequestration of soluble tubulin into an assembly-incompetent T2S complex (two α:β tubulin dimers per molecule of stathmin), and by increasing the switching frequency (catastrophe frequency) from growth to shortening at plus and minus ends by binding directly to the microtubules. Phosphorylation of stathmin on one or more of its four serine residues (Ser16, Ser25, Ser38, and Ser63) reduces its microtubule-destabilizing activity. However, the effects of phosphorylation of the individual serine residues of stathmin on microtubule dynamic instability have not been investigated systematically. Here we analyzed the effects of stathmin singly phosphorylated at Ser16 or Ser63, and doubly phosphorylated at Ser25 and Ser38, on its ability to modulate microtubule dynamic instability at steady-state in vitro. Phosphorylation at either Ser16 or Ser63 strongly reduced or abolished the ability of stathmin to bind to and sequester soluble tubulin and its ability to act as a catastrophe factor by directly binding to the microtubules. In contrast, double phosphorylation of Ser25 and Ser38 did not affect the binding of stathmin to tubulin or microtubules or its catastrophe-promoting activity. Our results indicate that the effects of stathmin on dynamic instability are strongly but differently attenuated by phosphorylation at Ser16 and Ser63 and support the hypothesis that selective targeting by Ser16-specific or Ser63-specific kinases provides complimentary mechanisms for regulating microtubule function.Stathmin is an 18-kDa ubiquitously expressed microtubule-destabilizing phosphoprotein whose activity is modulated by phosphorylation of its four serine residues, Ser16, Ser25, Ser38, and Ser63 (17). Several classes of kinases have been identified that phosphorylate stathmin, including kinases associated with cell growth and differentiation such as members of the mitogen-activated protein kinase (MAPK)2 family, cAMP-dependent protein kinase (15, 811), and kinases associated with cell cycle regulation such as cyclin-dependent kinase 1 (3, 1214). Phosphorylation of stathmin is required for cell cycle progression through mitosis and for proper assembly/function of the mitotic spindle (3, 1316). Inhibition of stathmin phosphorylation produces strong mitotic phenotypes characterized by disassembly and disorganization of mitotic spindles and abnormal chromosome distributions (3, 1314).Stathmin is known to destabilize microtubules in two ways. One is by binding to soluble tubulin and forming a stable complex that cannot polymerize into microtubules, consisting of one molecule of stathmin and two molecules of tubulin (T2S complex) (1724). Addition of stathmin to microtubules in equilibrium with soluble tubulin results in sequestration of the tubulin and a reduction in the level of microtubule polymer (1718, 22, 2528). In addition to reducing the amount of assembled polymer, tubulin sequestration by stathmin has been shown to increase the switching frequency at microtubule plus ends from growth to shortening (called the catastrophe frequency) as the microtubules relax to a new steady state (17, 29). The second way is by binding directly to microtubules (2730). The direct binding of stathmin to microtubules increases the catastrophe frequency at both ends of the microtubules and considerably more strongly at minus ends than at plus ends (27). Consistent with its strong catastrophe-promoting activity at minus ends, stathmin increases the treadmilling rate of steady-state microtubules in vitro (27). These results have led to the suggestion that stathmin might be an important cellular regulator of minus-end microtubule dynamics (27).Phosphorylation of stathmin diminishes its ability to regulate microtubule polymerization (3, 14, 2526). Phosphorylation of Ser16 or Ser63 appears to be more critical than phosphorylation of Ser25 and Ser38 for the ability of stathmin to bind to soluble tubulin and to inhibit microtubule assembly in vitro (3, 25). Inhibition of stathmin phosphorylation induces defects in spindle assembly and organization (3, 14) suggesting that not only soluble tubulin-microtubule levels are regulated by phosphorylation of stathmin, but the dynamics of microtubules could also be regulated in a phosphorylation-dependent manner.It is not known how phosphorylation at any of the four serine residues of stathmin affects its ability to regulate microtubule dynamics and, specifically, its ability to increase the catastrophe frequency at plus and minus ends due to its direct interaction with microtubules. Thus, we determined the effects of stathmin individually phosphorylated at either Ser16 or Ser63 and doubly phosphorylated at both Ser25 and Ser38 on dynamic instability at plus and minus ends in vitro at microtubule polymer steady state and physiological pH (pH 7.2). We find that phosphorylation of Ser16 strongly reduces the direct catastrophe-promoting activity of stathmin at plus ends and abolishes it at minus ends, whereas phosphorylation of Ser63 abolishes the activity at both ends. The effects of phosphorylation of individual serines correlated well with stathmin''s reduced abilities to form stable T2S complexes, to inhibit microtubule polymerization, and to bind to microtubules. In contrast, double phosphorylation of Ser25 and Ser38 did not alter the ability of stathmin to modulate dynamic instability at the microtubule ends, its ability to form a stable T2S complex, or its ability to bind to microtubules. The data further support the hypotheses that phosphorylation of stathmin on either Ser16 or Ser63 plays a critical role in regulating microtubule polymerization and dynamics in cells.  相似文献   

10.
11.
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   

12.
13.
14.
15.
16.
17.
18.
19.
20.
SLC26A7 (human)/Slc26a7 (mouse) is a recently identified chloride-base exchanger and/or chloride transporter that is expressed on the basolateral membrane of acid-secreting cells in the renal outer medullary collecting duct (OMCD) and in gastric parietal cells. Here, we show that mice with genetic deletion of Slc26a7 expression develop distal renal tubular acidosis, as manifested by metabolic acidosis and alkaline urine pH. In the kidney, basolateral Cl/HCO3 exchange activity in acid-secreting intercalated cells in the OMCD was significantly decreased in hypertonic medium (a normal milieu for the medulla) but was reduced only mildly in isotonic medium. Changing from a hypertonic to isotonic medium (relative hypotonicity) decreased the membrane abundance of Slc26a7 in kidney cells in vivo and in vitro. In the stomach, stimulated acid secretion was significantly impaired in isolated gastric mucosa and in the intact organ. We propose that SLC26A7 dysfunction should be investigated as a potential cause of unexplained distal renal tubular acidosis or decreased gastric acid secretion in humans.The collecting duct segment of the distal kidney nephron plays a major role in systemic acid base homeostasis by acid secretion and bicarbonate absorption. The acid secretion occurs via H+-ATPase and H-K-ATPase into the lumen and bicarbonate is absorbed via basolateral Cl/HCO3 exchangers (14). The tubules, which are located within the outer medullary region of the kidney collecting duct (OMCD),2 have the highest rate of acid secretion among the distal tubule segments and are therefore essential to the maintenance of acid base balance (2).The gastric parietal cell is the site of generation of acid and bicarbonate through the action of cytosolic carbonic anhydrase II (5, 6). The intracellular acid is secreted into the lumen via gastric H-K-ATPase, which works in conjunction with a chloride channel and a K+ recycling pathway (710). The intracellular bicarbonate is transported to the blood via basolateral Cl/HCO3 exchangers (1114).SLC26 (human)/Slc26 (mouse) isoforms are members of a conserved family of anion transporters that display tissue-specific patterns of expression in epithelial cells (1524). Several SLC26 members can function as chloride/bicarbonate exchangers. These include SLC26A3 (DRA), SLC26A4 (pendrin), SLC26A6 (PAT1 or CFEX), SLC26A7, and SLC26A9 (2531). SLC26A7 and SLC26A9 can also function as chloride channels (3234).SLC26A7/Slc26a7 is predominantly expressed in the kidney and stomach (28, 29). In the kidney, Slc26a7 co-localizes with AE1, a well-known Cl/HCO3 exchanger, on the basolateral membrane of (acid-secreting) A-intercalated cells in OMCD cells (29, 35, 36) (supplemental Fig. 1). In the stomach, Slc26a7 co-localizes with AE2, a major Cl/HCO3 exchanger, on the basolateral membrane of acid secreting parietal cells (28). To address the physiological function of Slc26a7 in the intact mouse, we have generated Slc26a7 ko mice. We report here that Slc26a7 ko mice exhibit distal renal tubular acidosis and impaired gastric acidification in the absence of morphological abnormalities in kidney or stomach.  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