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.
Calcium-binding protein 1 (CaBP1), a neuron-specific member of the calmodulin (CaM) superfamily, modulates Ca2+-dependent activity of inositol 1,4,5-trisphosphate receptors (InsP3Rs). Here we present NMR structures of CaBP1 in both Mg2+-bound and Ca2+-bound states and their structural interaction with InsP3Rs. CaBP1 contains four EF-hands in two separate domains. The N-domain consists of EF1 and EF2 in a closed conformation with Mg2+ bound at EF1. The C-domain binds Ca2+ at EF3 and EF4, and exhibits a Ca2+-induced closed to open transition like that of CaM. The Ca2+-bound C-domain contains exposed hydrophobic residues (Leu132, His134, Ile141, Ile144, and Val148) that may account for selective binding to InsP3Rs. Isothermal titration calorimetry analysis reveals a Ca2+-induced binding of the CaBP1 C-domain to the N-terminal region of InsP3R (residues 1-587), whereas CaM and the CaBP1 N-domain did not show appreciable binding. CaBP1 binding to InsP3Rs requires both the suppressor and ligand-binding core domains, but has no effect on InsP3 binding to the receptor. We propose that CaBP1 may regulate Ca2+-dependent activity of InsP3Rs by promoting structural contacts between the suppressor and core domains.Calcium ion (Ca2+) in the cell functions as an important messenger that controls neurotransmitter release, gene expression, muscle contraction, apoptosis, and disease processes (1). Receptor stimulation in neurons promotes large increases in intracellular Ca2+ levels controlled by Ca2+ release from intracellular stores through InsP3Rs (2). The neuronal type-1 receptor (InsP3R1)2 is positively and negatively regulated by cytosolic Ca2+ (3-6), important for the generation of repetitive Ca2+ transients known as Ca2+ spikes and waves (1). Ca2+-dependent activation of InsP3R1 contributes to the fast rising phase of Ca2+ signaling known as Ca2+-induced Ca2+ release (7). Ca2+-induced inhibition of InsP3R1, triggered at higher cytosolic Ca2+ levels, coordinates the temporal decay of Ca2+ transients (6). The mechanism of Ca2+-dependent regulation of InsP3Rs is complex (8, 9), and involves direct Ca2+ binding sites (5, 10) as well as remote sensing by extrinsic Ca2+-binding proteins such as CaM (11, 12), CaBP1 (13, 14), CIB1 (15), and NCS-1 (16).Neuronal Ca2+-binding proteins (CaBP1-5 (17)) represent a new sub-branch of the CaM superfamily (18) that regulate various Ca2+ channel targets. Multiple splice variants and isoforms of CaBPs are localized in different neuronal cell types (19-21) and perform specialized roles in signal transduction. CaBP1, also termed caldendrin (22), has been shown to modulate the Ca2+-sensitive activity of InsP3Rs (13, 14). CaBP1 also regulates P/Q-type voltage-gated Ca2+ channels (23), L-type channels (24), and the transient receptor potential channel, TRPC5 (25). CaBP4 regulates Ca2+-dependent inhibition of L-type channels in the retina and may be genetically linked to retinal degeneration (26). Thus, the CaBP proteins are receiving increased attention as a family of Ca2+ sensors that control a variety of Ca2+ channel targets implicated in neuronal degenerative diseases.CaBP proteins contain four EF-hands, similar in sequence to those found in CaM and troponin C (18) (Fig. 1). By analogy to CaM (27), the four EF-hands are grouped into two domains connected by a central linker that is four residues longer in CaBPs than in CaM. In contrast to CaM, the CaBPs contain non-conserved amino acids within the N-terminal region that may confer target specificity. Another distinguishing property of CaBPs is that the second EF-hand lacks critical residues required for high affinity Ca2+ binding (17). CaBP1 binds Ca2+ only at EF3 and EF4, whereas it binds Mg2+ at EF1 that may serve a functional role (28). Indeed, changes in cytosolic Mg2+ levels have been detected in cortical neurons after treatment with neurotransmitter (29). Other neuronal Ca2+-binding proteins such as DREAM (30), CIB1 (31), and NCS-1 (32) also bind Mg2+ and exhibit Mg2+-induced physiological effects. Mg2+ binding in each of these proteins helps stabilize their Ca2+-free state to interact with signaling targets.Open in a separate windowFIGURE 1.Amino acid sequence alignment of human CaBP1 with CaM. Secondary structural elements (α-helices and β-strands) were derived from NMR analysis. The four EF-hands (EF1, EF2, EF3, and EF4) are highlighted green, red, cyan, and yellow. Residues in the 12-residue Ca2+-binding loops are underlined and chelating residues are highlighted bold. Non-conserved residues in the hydrophobic patch are colored red.Despite extensive studies on CaBP1, little is known about its structure and target binding properties, and regulation of InsP3Rs by CaBP1 is somewhat controversial and not well understood. Here, we present the NMR solution structures of both Mg2+-bound and Ca2+-bound conformational states of CaBP1 and their structural interactions with InsP3R1. These CaBP1 structures reveal important Ca2+-induced structural changes that control its binding to InsP3R1. Our target binding analysis demonstrates that the C-domain of CaBP1 exhibits Ca2+-induced binding to the N-terminal cytosolic region of InsP3R1. We propose that CaBP1 may regulate Ca2+-dependent channel activity in InsP3Rs by promoting a structural interaction between the N-terminal suppressor and ligand-binding core domains that modulates Ca2+-dependent channel gating (8, 33, 34).  相似文献   

2.
Calmodulin (CaM) is a highly conserved intracellular calcium sensor. In plants, CaM also appears to be present in the apoplasm, and application of exogenous CaM has been shown to influence a number of physiological functions as a polypeptide signal; however, the existence and localization of its corresponding apoplasmic binding sites remain controversial. To identify the site(s) of action, a CaM-conjugated quantum dot (QD) system was employed for single molecule level detection at the surface of plant cells. Using this approach, we show that QD-CaM binds selectively to sites on the outer surface of the plasma membrane, which was further confirmed by high resolution transmission electron microscopy. Measurements of Ca2+ fluxes across the plasma membrane, using ion-selective microelectrodes, demonstrated that exogenous CaM induces a net influx into protoplasts. Consistent with these flux studies, calcium-green-dextran and FRET experiments confirmed that applied CaM/QD-CaM elicited an increase in cytoplasmic Ca2+ levels. These results support the hypothesis that apoplasmic CaM can act as a signaling agent. These findings are discussed in terms of CaM acting as an apoplasmic peptide ligand to mediate transmembrane signaling in the plant kingdom.Calmodulin (CaM)2 is a conserved multifunctional calcium sensor that mediates intracellular Ca2+ signaling and regulates diverse cellular processes by interacting with calmodulin-binding proteins (13). Interestingly, in both animals and plants, CaM may also act as an extracellular agent to regulate physiological events (4). Consistent with this notion, extracellular CaM has been detected within the cell walls of a broad range of plant species (4, 5).Functional studies have established that exogenously applied CaM can stimulate the proliferation of suspension-cultured plant cells (6) as well as affect intracellular activities of heterotrimeric G proteins and phospholipases in protoplasts (7, 8). Based on these findings, it has been proposed that, in plants, extracellular CaM may function as a signaling agent involved in the regulation of cell growth and development (4). However, as a 17-kDa hydrophilic protein, exogenously applied CaM could well be retrieved from the apoplasmic space and then exert its effects on components within the cytoplasm. Evidence against this hypothesis was provided by studies with Arabidopsis thaliana suspension-cultured cells in which it was shown that 24 h of incubation in exogenous CaM did not result in protein uptake or degradation (4).To exert an effect from the apoplasm, it would seem logical to assume that a protein(s) within the plant plasma membrane would have a CaM-binding site exposed to the apoplasm. Although a number of studies have addressed the molecular mechanism(s) by which extracellular CaM might act as a signal (6, 9) and attempts have been made to identify extracellular CaM-binding proteins (4, 6), currently there is no direct evidence in support of the hypothesis that specific CaM-binding sites exist at the surface of plant cells.To address this question, a CaM-conjugated quantum dot (QD) system was employed for single molecule level detection (1013) at the surface of plant cells. These nanoparticles have several advantages over conventional fluorophores for light microscopic imaging, including their higher brightness and photostability (14, 15). In addition, because of their electron dense nature, QDs can be used for single labeling studies at the transmission electron microscope level (16, 17). Using this QD-CaM system, we demonstrate that QD-CaM binds selectively to sites on the outer surface of the plant plasma membrane. We also show by three independent methods that applied CaM can modulate Ca2+ fluxes across the plasma membrane, leading to alterations in cytoplasmic Ca2+ status. These findings support the hypothesis that, in plants, apoplasmic CaM can act as a signaling agent.  相似文献   

3.
We have investigated the possible biochemical basis for enhancements in NO production in endothelial cells that have been correlated with agonist- or shear stress-evoked phosphorylation at Ser-1179. We have found that a phosphomimetic substitution at Ser-1179 doubles maximal synthase activity, partially disinhibits cytochrome c reductase activity, and lowers the EC50(Ca2+) values for calmodulin binding and enzyme activation from the control values of 182 ± 2 and 422 ± 22 nm to 116 ± 2 and 300 ± 10 nm. These are similar to the effects of a phosphomimetic substitution at Ser-617 (Tran, Q. K., Leonard, J., Black, D. J., and Persechini, A. (2008) Biochemistry 47, 7557–7566). Although combining substitutions at Ser-617 and Ser-1179 has no additional effect on maximal synthase activity, cooperativity between the two substitutions completely disinhibits reductase activity and further reduces the EC50(Ca2+) values for calmodulin binding and enzyme activation to 77 ± 2 and 130 ± 5 nm. We have confirmed that specific Akt-catalyzed phosphorylation of Ser-617 and Ser-1179 and phosphomimetic substitutions at these positions have similar functional effects. Changes in the biochemical properties of eNOS produced by combined phosphorylation at Ser-617 and Ser-1179 are predicted to substantially increase synthase activity in cells at a typical basal free Ca2+ concentration of 50–100 nm.The nitric-oxide synthases catalyze formation of NO and l-citrulline from l-arginine and O2, with NADPH as the electron donor (1). The role of NO generated by endothelial nitricoxide synthase (eNOS)2 in the regulation of smooth muscle tone is well established and was the first of several physiological roles for this small molecule that have so far been identified (2). The nitric-oxide synthases are homodimers of 130–160-kDa subunits. Each subunit contains a reductase and oxygenase domain (1). A significant difference between the reductase domains in eNOS and nNOS and the homologous P450 reductases is the presence of inserts in these synthase isoforms that appear to maintain them in their inactive states (3, 4). A calmodulin (CaM)-binding domain is located in the linker that connects the reductase and oxygenase domains, and the endothelial and neuronal synthases both require Ca2+ and exogenous CaM for activity (5, 6). When CaM is bound, it somehow counteracts the effects of the autoinhibitory insert(s) in the reductase. The high resolution structure for the complex between (Ca2+)4-CaM and the isolated CaM-binding domain from eNOS indicates that the C-ter and N-ter lobes of CaM, which each contain a pair of Ca2+-binding sites, enfold the domain, as has been observed in several other such CaM-peptide complexes (7). Consistent with this structure, investigations of CaM-dependent activation of the neuronal synthase suggest that both CaM lobes must participate (8, 9).Bovine eNOS can be phosphorylated in endothelial cells at Ser-116, Thr-497, Ser-617, Ser-635, and Ser-1179 (1012). There are equivalent phosphorylation sites in the human enzyme (1012). Phosphorylation of the bovine enzyme at Thr-497, which is located in the CaM-binding domain, blocks CaM binding and enzyme activation (7, 11, 13, 14). Ser-116 can be basally phosphorylated in cells (10, 11, 13, 15), and dephosphorylation of this site has been correlated with increased NO production (13, 15). However, it has also been reported that a phosphomimetic substitution at this position has no effect on enzyme activity measured in vitro (13). Ser-1179 is phosphorylated in response to a variety of stimuli, and this has been reliably correlated with enhanced NO production in cells (10, 11). Indeed, NO production is elevated in transgenic endothelium expressing an eNOS mutant containing an S1179D substitution, but not in tissue expressing an S1179A mutant (16). Shear stress or insulin treatment is correlated with Akt-catalyzed phosphorylation of Ser-1179 in endothelial cells, and this is correlated with increased NO production in the absence of extracellular Ca2+ (1719). Akt-catalyzed phosphorylation or an S1179D substitution has also been correlated with increased synthase activity in cell extracts at low intracellular free [Ca2+] (17). Increased NO production has also been observed in cells expressing an eNOS mutant containing an S617D substitution, and physiological stimuli such as shear-stress, bradykinin, VEGF, and ATP appear to stimulate Akt-catalyzed phosphorylation of Ser-617 and Ser-1179 (12, 13, 20). Although S617D eNOS has been reported to have the same maximum activity in vitro as the wild type enzyme (20), in our hands an S617D substitution increases the maximal CaM-dependent synthase activity of purified mutant enzyme ∼2-fold, partially disinhibits reductase activity, and reduces the EC50(Ca2+) values for CaM binding and enzyme activation (21).In this report, we describe the effects of a phosphomimetic Asp substitution at Ser-1179 in eNOS on the Ca2+ dependence of CaM binding and CaM-dependent activation of reductase and synthase activities. We also describe the effects on these properties of combining this substitution with one at Ser-617. Finally, we demonstrate that Akt-catalyzed phosphorylation and Asp substitutions at Ser-617 and Ser-1179 have similar functional effects. Our results suggest that phosphorylation of eNOS at Ser-617 and Ser-1179 can substantially increase synthase activity in cells at a typical basal free Ca2+ concentration of 50–100 nm, while single phosphorylations at these sites produce smaller activity increases, and can do so only at higher free Ca2+ concentrations.  相似文献   

4.
The Ca2+ release-activated Ca2+ channel is a principal regulator of intracellular Ca2+ rise, which conducts various biological functions, including immune responses. This channel, involved in store-operated Ca2+ influx, is believed to be composed of at least two major components. Orai1 has a putative channel pore and locates in the plasma membrane, and STIM1 is a sensor for luminal Ca2+ store depletion in the endoplasmic reticulum membrane. Here we have purified the FLAG-fused Orai1 protein, determined its tetrameric stoichiometry, and reconstructed its three-dimensional structure at 21-Å resolution from 3681 automatically selected particle images, taken with an electron microscope. This first structural depiction of a member of the Orai family shows an elongated teardrop-shape 150Å in height and 95Å in width. Antibody decoration and volume estimation from the amino acid sequence indicate that the widest transmembrane domain is located between the round extracellular domain and the tapered cytoplasmic domain. The cytoplasmic length of 100Å is sufficient for direct association with STIM1. Orifices close to the extracellular and intracellular membrane surfaces of Orai1 seem to connect outside the molecule to large internal cavities.Ca2+ is an intracellular second messenger that plays important roles in various physiological functions such as immune response, muscle contraction, neurotransmitter release, and cell proliferation. Intracellular Ca2+ is mainly stored in the endoplasmic reticulum (ER).2 This ER system is distributed through the cytoplasm from around the nucleus to the cell periphery close to the plasma membrane. In non-excitable cells, the ER releases Ca2+ through the inositol 1,4,5-trisphosphate (IP3) receptor channel in response to various signals, and the Ca2+ store is depleted. Depletion of Ca2+ then induces Ca2+ influx from outside the cell to help in refilling the Ca2+ stores and to continue Ca2+ rise for several minutes in the cytoplasm (1, 2). This Ca2+ influx was first proposed by Putney (3) and was named store-operated Ca2+ influx. In the immune system, store-operated Ca2+ influx is mainly mediated by the Ca2+ release-activated Ca2+ (CRAC) current, which is a highly Ca2+-selective inwardly rectified current with low conductance (4, 5). Pathologically, the loss of CRAC current in T cells causes severe combined immunodeficiency (6) where many Ca2+ signal-dependent gene expressions, including cytokines, are interrupted (7). Therefore, CRAC current is necessary for T cell functions.Recently, Orai1 (also called CRACM1) and STIM1 have been physiologically characterized as essential components of the CRAC channel (812). They are separately located in the plasma membrane and in the ER membrane; co-expression of these proteins presents heterologous CRAC-like currents in various types of cells (10, 1315). Both of them are shown to be expressed ubiquitously in various tissues (1618). STIM1 senses Ca2+ depletion in the ER through its EF hand motif (19) and transmits a signal to Orai1 in the plasma membrane. Although Orai1 is proposed as a regulatory component for some transient receptor potential canonical channels (20, 21), it is believed from the mutation analyses to be the pore-forming subunit of the CRAC channel (8, 2224). In the steady state, both Orai1 and STIM1 molecules are dispersed in each membrane. When store depletion occurs, STIM1 proteins gather into clusters to form puncta in the ER membrane near the plasma membrane (11, 19). These clusters then trigger the clustering of Orai1 in the plasma membrane sites opposite the puncta (25, 26), and CRAC channels are activated (27).Orai1 has two homologous genes, Orai2 and Orai3 (8). They form the Orai family and have in common the four transmembrane (TM) segments with relatively large N and C termini. These termini are demonstrated to be in the cytoplasm, because both N- and C-terminally introduced tags are immunologically detected only in the membrane-permeabilized cells (8, 9). The subunit stoichiometry of Orai1 is as yet controversial: it is believed to be an oligomer, presumably a dimer or tetramer even in the steady state (16, 2830).Despite the accumulation of biochemical and electrophysiological data, structural information about Orai1 is limited due to difficulties in purification and crystallization. In this study, we have purified Orai1 in its tetrameric form and have reconstructed the three-dimensional structure from negatively stained electron microscopic (EM) images.  相似文献   

5.
6.
Oxygen and glucose deprivation (OGD) induces delayed cell death in hippocampal CA1 neurons via Ca2+/Zn2+-permeable, GluR2-lacking AMPA receptors (AMPARs). Following OGD, synaptic AMPAR currents in hippocampal neurons show marked inward rectification and increased sensitivity to channel blockers selective for GluR2-lacking AMPARs. This occurs via two mechanisms: a delayed down-regulation of GluR2 mRNA expression and a rapid internalization of GluR2-containing AMPARs during the OGD insult, which are replaced by GluR2-lacking receptors. The mechanisms that underlie this rapid change in subunit composition are unknown. Here, we demonstrate that this trafficking event shares features in common with events that mediate long term depression and long term potentiation and is initiated by the activation of N-methyl-d-aspartic acid receptors. Using biochemical and electrophysiological approaches, we show that peptides that interfere with PICK1 PDZ domain interactions block the OGD-induced switch in subunit composition, implicating PICK1 in restricting GluR2 from synapses during OGD. Furthermore, we show that GluR2-lacking AMPARs that arise at synapses during OGD as a result of PICK1 PDZ interactions are involved in OGD-induced delayed cell death. This work demonstrates that PICK1 plays a crucial role in the response to OGD that results in altered synaptic transmission and neuronal death and has implications for our understanding of the molecular mechanisms that underlie cell death during stroke.Oxygen and glucose deprivation (OGD)3 associated with transient global ischemia induces delayed cell death, particularly in hippocampal CA1 pyramidal cells (13), a phenomenon that involves Ca2+/Zn2+-permeable, GluR2-lacking AMPARs (4). AMPARs are heteromeric complexes of subunits GluR1–4 (5), and most AMPARs in the hippocampus contain GluR2, which renders them calcium-impermeable and results in a marked inward rectification in their current-voltage relationship (68). Ischemia induces a delayed down-regulation of GluR2 mRNA and protein expression (4, 911), resulting in enhanced AMPAR-mediated Ca2+ and Zn2+ influx into CA1 neurons (10, 12). In these neurons, AMPAR-mediated postsynaptic currents (EPSCs) show marked inward rectification 1–2 days following ischemia and increased sensitivity to 1-naphthyl acetyl spermine (NASPM), a channel blocker selective for GluR2-lacking AMPARs (1316). Blockade of these channels at 9–40 h following ischemia is neuroprotective, indicating a crucial role for Ca2+-permeable AMPARs in ischemic cell death (16).In addition to delayed changes in AMPAR subunit composition as a result of altered mRNA expression, it was recently reported that Ca2+-permable, GluR2-lacking AMPARs are targeted to synaptic sites via membrane trafficking at much earlier times during OGD (17). This subunit rearrangement involves endocytosis of AMPARs containing GluR2 complexed with GluR1/3, followed by exocytosis of GluR2-lacking receptors containing GluR1/3 (17). However, the molecular mechanisms behind this trafficking event are unknown, and furthermore, it is not known whether these trafficking-mediated changes in AMPAR subunit composition contribute to delayed cell death.AMPAR trafficking is a well studied phenomenon because of its crucial involvement in long term depression (LTD) and long term potentiation (LTP), activity-dependent forms of synaptic plasticity thought to underlie learning and memory. AMPAR endocytosis, exocytosis, and more recently subunit-switching events (brought about by trafficking that involves endo/exocytosis) are central to the necessary changes in synaptic receptor complement (7, 1820). It is possible that similar mechanisms regulate AMPAR trafficking during OGD.PICK1 is a PDZ and BAR (Bin-amphiphysin-Rus) domain-containing protein that binds, via the PDZ domain, to a number of membrane proteins including AMPAR subunits GluR2/3. This interaction is required for AMPAR internalization from the synaptic plasma membrane in response to Ca2+ influx via NMDAR activation in hippocampal neurons (2123). This process is the major mechanism that underlies the reduction in synaptic strength in LTD. Furthermore, PICK1-mediated trafficking has recently emerged as a mechanism that regulates the GluR2 content of synaptic receptors, which in turn determines their Ca2+ permeability (7, 20). This is likely to be of profound importance in both plasticity and pathological mechanisms. Importantly, PICK1 overexpression has been shown to induce a shift in synaptic AMPAR subunit composition in hippocampal CA1 neurons, resulting in inwardly rectifying AMPAR EPSCs via reduced surface GluR2 and no change in GluR1 (24). This suggests that PICK1 may mediate the rapid switch in subunit composition occurring during OGD (17). Here, we demonstrate that the OGD-induced switch in AMPAR subunit composition is dependent on PICK1 PDZ interactions, and importantly, that this early trafficking event that occurs during OGD contributes to the signaling that results in delayed neuronal death.  相似文献   

7.
The troponin complex on the thin filament plays a crucial role in the regulation of muscle contraction. However, the precise location of troponin relative to actin and tropomyosin remains uncertain. We have developed a method of reconstructing thin filaments using single particle analysis that does not impose the helical symmetry of actin and is independent of a starting model. We present a single particle three-dimensional reconstruction of the thin filament. Atomic models of the F-actin filament were fitted into the electron density maps and troponin and tropomyosin located. The structure provides evidence that the globular head region of troponin labels the two strands of actin with a 27.5-Å axial stagger. The density attributed to troponin appears tapered with the widest point toward the barbed end. This leads us to interpret the polarity of the troponin complex in the thin filament as reversed with respect to the widely accepted model.Regulation of actin filament function is a fundamental biological process with implications ranging from cell migration to muscle contraction. Skeletal and cardiac muscle thin filaments consist of actin and the regulatory proteins troponin and tropomyosin. Contraction is initiated by release of Ca2+ into the sarcomere and the consequent binding of Ca2+ to regulatory sites on troponin. Troponin is believed to undergo a conformational change leading to an azimuthal movement of tropomyosin, which allows myosin heads to interact with actin, hydrolyze ATP, and generate force. The molecular basis by which troponin acts to regulate muscle contraction is only partly understood. It is essential that the structure of troponin in the thin filament at high and low Ca2+ is determined to properly understand the mechanism of regulation.The basic structure of the thin filament was described by Ebashi in 1972 (1). In this structure each tropomyosin molecule covers seven actin monomers, and there is a 27.5-Å stagger between troponin molecules. The 7-Å tropomyosin structure (2), the atomic model of F-actin (3), and the troponin “core domain” (4) have recently been used to generate atomic models of the thin filament in low and high Ca2+ states (5). While the position of troponin in these models was constrained by known distance measurements between filament components, the exact arrangement of the complex on the filament has not been determined a priori. Although recently published crystal structures of partial troponin complexes (4, 6) have provided valuable insights into the arrangement of the globular head or core domain, the complex in its entirety has not been crystallized.Troponin is believed to consist of a globular core domain with an extended tail (7). The globular core contains the Ca2+-binding subunit (TnC),2 the inhibitory subunit (TnI), and the C-terminal part (residues 156–262) of the tropomyosin-binding subunit (TnT). The extended tail consists of the N-terminal part of TnT (residues 1–155). A structural rearrangement associated with Ca2+ dissociation from the troponin core has been observed (4) such that the helix connecting the two domains of TnC collapses, releasing the TnI inhibitory segment. It is postulated that the TnI inhibitory segment then becomes able to bind actin, in so doing biasing tropomyosin (8). To understand properly how Ca2+ binding to TnC leads to movement of tropomyosin, it is necessary to determine a high resolution structure of troponin attached to the thin filament, allowing unambiguous docking of the available crystal structures and direct observation of any changes at a molecular level caused by Ca2+ binding.Direct visualization of the thin filament is possible using electron microscopy. Tropomyosin strands have been resolved in the low and high Ca2+ states confirming the movement of tropomyosin and the steric blocking model (9, 10). Until recently the actin helical repeat has been imposed in the majority of reconstructions of the thin filament causing artifacts. Helical averaging using the actin repeat spreads troponin density over every actin monomer, which prevents the detailed position and shape of the troponin complex from being found (11). It is possible to avoid this effect by applying a single particle approach. Individual filament images are divided into segments and each segment treated as a particle. Three-dimensional reconstruction may then be carried out by single particle techniques of alignment, classification (12, 13), Euler angle assignment (1416) and exact filter back-projection (17, 18).Two forms of single particle analysis have emerged: helical single particle analysis (19), where the determined helical symmetry is applied to the final reconstruction, and non-helical single particle analysis, which treats the complex as a truly asymmetric particle. Helical single particle analysis has been used to successfully reconstruct a myosin containing invertebrate thick filament to a resolution of 25 Å (20), and non-helical single particle analysis has been applied to the vertebrate skeletal muscle thick filament allowing azimuthal perturbations of the myosin heads to be observed (21).Model-based single particle image processing methods have recently been applied to the structural analysis of the vertebrate (5, 22, 23) and the insect thin filament (24). We have deliberately avoided starting with a model and any potential model bias by using a reference-free alignment procedure. The adaptation of conventional procedures and their application to the structural study of the muscle thin filament has been documented (25).  相似文献   

8.
9.
The molecular mechanisms underlying synaptic exocytosis in the hair cell, the auditory and vestibular receptor cell, are not well understood. Otoferlin, a C2 domain-containing Ca2+-binding protein, has been implicated as having a role in vesicular release. Mutations in the OTOF gene cause nonsyndromic deafness in humans, and OTOF knock-out mice are deaf. In the present study, we generated otoferlin fusion proteins containing two of the same amino acid substitutions detected in DFNB9 patients (P1825A in C2F and L1011P in C2D). The native otoferlin C2F domain bound syntaxin 1A and SNAP-25 in a Ca2+-dependent manner (with optimal 61 μm free Ca2+ required for binding). These interactions were greatly diminished for C2F with the P1825A mutation, possibly because of a reduction in tertiary structural change, induced by Ca2+, for the mutated C2F compared with the native C2F. The otoferlin C2D domain also bound syntaxin 1A, but with weaker affinity (Kd = 1.7 × 10–5 m) than for the C2F interaction (Kd = 2.6 × 10–9 m). In contrast, it was the otoferlin C2D domain that bound the Cav1.3 II-III loop, in a Ca2+-dependent manner. The L1011P mutation in C2D rendered this binding insensitive to Ca2+ and considerably diminished. Overall, we demonstrated that otoferlin interacts with two main target-SNARE proteins of the hair-cell synaptic complex, syntaxin 1A and SNAP-25, as well as the calcium channel, with the otoferlin C2F and C2D domains of central importance for binding. Because mutations in the otoferlin C2 domains that cause deafness in humans impair the ability of otoferlin to bind syntaxin, SNAP-25, and the Cav1.3 calcium channel, it is these interactions that may mediate regulation by otoferlin of hair cell synaptic exocytosis critical to inner ear hair cell function.Calcium is a key regulator of synaptic vesicle fusion (reviewed in Ref. 1). In mechanosensory hair cells, calcium microdomains (2) and possibly nanodomains (3) are formed when voltage-gated calcium channels open upon depolarization. Calcium at these sites is thought to activate protein interactions, leading to vesicle fusion. Some of the key players in this process are the target-SNARE2 proteins, syntaxin 1A and SNAP-25, and the vesicle-SNARE, synaptobrevin (4). Vesicle-SNARE synaptotagmin 1 plays a crucial role as a calcium sensor at the neuronal synapse, modulating calcium channels and vesicle release by a Ca2+-dependent interaction with other SNARE proteins in the presence of lipid molecules (46). However, in vertebrate mechanosensory hair cells, synaptotagmin 1 is not detected (7). Instead, fast neurotransmitter release in auditory and vestibular hair cells, facilitated largely by an L-type voltagegated calcium channel, Cav1.3 (8, 9), is thought to be modulated by a newly discovered protein, otoferlin, acting as the Ca2+ sensor and vesicle-binding protein. When mutated, otoferlin causes DFNB9 nonsyndromic deafness (10). Gene sequences of different deaf families show that the OTOF gene can undergo mutation at multiple locations (1113). Recently, it has been demonstrated that otoferlin is necessary for synaptic exocytosis from hair cells (14). Further, an engineered mutation in the C2B domain of otoferlin has been shown to cause deafness in mice (15). However, the precise function of otoferlin as a synaptic protein is not well understood.Specific mutations in the otoferlin C2F (P1825A) or C2D (L1011P) domains in humans have been documented to cause DFNB9 deafness (11, 12). Previous studies suggested that a region of otoferlin containing all three C2 domains, D, E, and F, binds directly to the t-SNARE molecules syntaxin 1A and SNAP-25 in response to an increase in Ca2+ concentration (14). However, it is not understood how a single amino acid substitution in one domain of otoferlin, such as C2F (11) or C2D (12), might independently lead to deafness. Here, we examine the role of otoferlin as a Ca2+ sensor as well as a facilitator of vesicle fusion, as indicated by protein-protein interactions and their [Ca2+] dependence.  相似文献   

10.
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   

11.
Galectin-1 (Gal-1) regulates leukocyte turnover by inducing the cell surface exposure of phosphatidylserine (PS), a ligand that targets cells for phagocytic removal, in the absence of apoptosis. Gal-1 monomer-dimer equilibrium appears to modulate Gal-1-induced PS exposure, although the mechanism underlying this regulation remains unclear. Here we show that monomer-dimer equilibrium regulates Gal-1 sensitivity to oxidation. A mutant form of Gal-1, containing C2S and V5D mutations (mGal-1), exhibits impaired dimerization and fails to induce cell surface PS exposure while retaining the ability to recognize carbohydrates and signal Ca2+ flux in leukocytes. mGal-1 also displayed enhanced sensitivity to oxidation, whereas ligand, which partially protected Gal-1 from oxidation, enhanced Gal-1 dimerization. Continual incubation of leukocytes with Gal-1 resulted in gradual oxidative inactivation with concomitant loss of cell surface PS, whereas rapid oxidation prevented mGal-1 from inducing PS exposure. Stabilization of Gal-1 or mGal-1 with iodoacetamide fully protected Gal-1 and mGal-1 from oxidation. Alkylation-induced stabilization allowed Gal-1 to signal sustained PS exposure in leukocytes and mGal-1 to signal both Ca2+ flux and PS exposure. Taken together, these results demonstrate that monomer-dimer equilibrium regulates Gal-1 sensitivity to oxidative inactivation and provides a mechanism whereby ligand partially protects Gal-1 from oxidation.Immunological homeostasis relies on efficient contraction of activated leukocytes following an inflammatory episode. Several factors, including members of the galectin and tumor necrosis factor families (1, 2), regulate leukocyte turnover by inducing apoptotic cell death. In contrast, several galectin family members, in particular galectin-1 (Gal-1),2 uniquely regulate neutrophil turnover by inducing phosphatidylserine (PS) exposure, which normally sensitizes apoptotic cells to phagocytic removal (3, 4), independent of apoptosis, a process recently termed preaparesis (5).Previous studies suggested that dimerization may be required for Gal-1-induced PS exposure, as a mutant form of Gal-1 (mGal-1) containing two point mutations within the dimer interface, C2S and V5D (C2S,V5D), displays impaired Gal-1 dimerization and fails to induce PS exposure (6). However, the manner in which monomer-dimer equilibrium regulates Gal-1 signaling remains unclear. Previous studies suggest that dimerization may be required for efficient cross-linking of functional receptors or the formation of signaling lattices (79). Consistent with this, monomeric mutants of several other galectins fail to induce PS exposure or signal leukocytes (4, 8). Gal-1 signaling of PS exposure requires initial signaling events, such as mobilization of intracellular Ca2+ followed by sustained receptor engagement (10). Although mGal-1 fails to induce PS exposure (6), whether mGal-1 can induce these initial signaling events remains unknown (10).In addition to directly regulating signaling, monomer-dimer equilibrium may also regulate other aspects of Gal-1 function. Unlike many other proteins involved in the regulation of immunity, Gal-1 displays unique sensitivity to oxidative inactivation (1115). Although engagement of ligand partially protects Gal-1 from oxidation (15), the impact of Gal-1 oxidation on signaling remains enigmatic. During oxidation, Gal-1 forms three distinct intramolecular disulfide bridges that facilitate profound conformational changes that preclude ligand binding and Gal-1 dimerization (1214), suggesting that monomerdimer equilibrium may also regulate Gal-1 sensitivity to oxidative inactivation.Previous studies utilized dithiothreitol (DTT) in treatment conditions to protect Gal-1 from oxidative inactivation (16, 17). Indeed, failure to include DTT precluded Gal-1-induced death in T cells (3, 18), suggesting that Gal-1 undergoes rapid oxidation in vivo in the absence of reducing conditions. However, DTT itself can induce apoptosis in leukocytes (19), leaving questions regarding the impact of Gal-1 oxidation on these signaling events. In contrast, recent studies utilizing iodoacetamide-alkylated Gal-1 (iGal-1), previously shown to protect Gal-1 from oxidative inactivation (2029), demonstrated that DTT actually primes cells to become sensitive to Gal-1-induced apoptosis regardless of Gal-1 sensitivity to oxidation (5).As the engagement of leukocyte ligands requires glycan recognition and oxidation precludes this binding (11, 15), understanding the impact of oxidation on Gal-1 signals will facilitate a greater appreciation of the factors that govern Gal-1 oxidation and therefore function. Our results demonstrate that Gal-1 monomer-dimer equilibrium provides a key regulatory point controlling both Gal-1 sensitivity to oxidation and its ability to signal PS exposure in leukocytes. These results provide novel insights into Gal-1 function and explain at a biochemical level the mechanisms regulating Gal-1 oxidative inactivation and signaling.  相似文献   

12.
Zinc is hypothesized to be co-released with glutamate at synapses of the central nervous system. Zinc binds to NR1/NR2A N-methyl-d-aspartate (NMDA) receptors with high affinity and inhibits NMDAR function in a voltage-independent manner. The serine protease plasmin can cleave a number of substrates, including protease-activated receptors, and may play an important role in several disorders of the central nervous system, including ischemia and spinal cord injury. Here, we demonstrate that plasmin can cleave the native NR2A amino-terminal domain (NR2AATD), removing the functional high affinity Zn2+ binding site. Plasmin also cleaves recombinant NR2AATD at lysine 317 (Lys317), thereby producing a ∼40-kDa fragment, consistent with plasmin-induced NR2A cleavage fragments observed in rat brain membrane preparations. A homology model of the NR2AATD predicts that Lys317 is near the surface of the protein and is accessible to plasmin. Recombinant expression of NR2A with an amino-terminal deletion at Lys317 is functional and Zn2+ insensitive. Whole cell voltage-clamp recordings show that Zn2+ inhibition of agonist-evoked NMDA receptor currents of NR1/NR2A-transfected HEK 293 cells and cultured cortical neurons is significantly reduced by plasmin treatment. Mutating the plasmin cleavage site Lys317 on NR2A to alanine blocks the effect of plasmin on Zn2+ inhibition. The relief of Zn2+ inhibition by plasmin occurs in PAR1-/- cortical neurons and thus is independent of interaction with protease-activated receptors. These results suggest that plasmin can directly interact with NMDA receptors, and plasmin may increase NMDA receptor responses through disruption or removal of the amino-terminal domain and relief of Zn2+ inhibition.N-Methyl-d-aspartate (NMDA)2 receptors are one of three types of ionotropic glutamate receptors that play critical roles in excitatory neurotransmission, synaptic plasticity, and neuronal death (13). NMDA receptors are comprised of glycine-binding NR1 subunits in combination with at least one type of glutamate-binding NR2 subunit (1, 4). Each subunit contains three transmembrane domains, one cytoplasmic re-entrant membrane loop, one bi-lobed domain that forms the ligand binding site, and one bi-lobed amino-terminal domain (ATD), thought to share structural homology to periplasmic amino acid-binding proteins (46). Activation of NMDA receptors requires combined stimulation by glutamate and the co-agonist glycine in addition to membrane depolarization to overcome voltage-dependent Mg2+ block of the ion channel (7). The activity of NMDA receptors is negatively modulated by a variety of extracellular ions, including Mg2+, polyamines, protons, and Zn2+ ions, which can exert tonic inhibition under physiological conditions (1, 4). Several extracellular modulators such as Zn2+ and ifenprodil are thought to act at the ATD of the NMDA receptor (814).Zinc is a transition metal that plays key roles in both catalytic and structural capacities in all mammalian cells (15). Zinc is required for normal growth and survival of cells. In addition, neuronal death in hypoxia-ischemia and epilepsy has been associated with Zn2+ (1618). Abnormal metabolism of zinc may contribute to induction of cytotoxicity in neurodegenerative diseases, such as Alzheimer''s disease, Parkinson''s disease, and amyotrophic lateral sclerosis (19). Zinc is co-released with glutamate at excitatory presynaptic terminals and inhibits native NMDA receptor activation (20, 21). Zn2+ inhibits NMDA receptor function through a dual mechanism, which includes voltage-dependent block and voltage-independent inhibition (2224). Voltage-independent Zn2+ inhibition at low nanomolar concentrations (IC50, 20 nm) is observed for NR2A-containing NMDA receptors (2528). Evidence has accumulated that the amino-terminal domain of the NR2A subunit controls high-affinity Zn2+ inhibition of NMDA receptors, and several histidine residues in this region may constitute part of an NR2A-specific Zn2+ binding site (8, 9, 11, 12). For the NR2A subunit, several lines of evidence suggest that Zn2+ acts by enhancing proton inhibition (8, 11, 29, 30).Serine proteases present in the circulation, mast cells, and elsewhere signal directly to cells by cleaving protease-activated receptors (PARs), members of a subfamily of G-protein-coupled receptors. Cleavage exposes a tethered ligand domain that binds to and activates the cleaved receptors (31, 32). Protease receptor activation has been studied extensively in relation to coagulation and thrombolysis (33). In addition to their circulation in the bloodstream, some serine proteases and PARs are expressed in the central nervous system, and have been suggested to play roles in physiological conditions (e.g. long-term potentiation or memory) and pathophysiological states such as glial scarring, edema, seizure, and neuronal death (31, 3436).Functional interactions between proteases and NMDA receptors have previously been suggested. Earlier studies reported that the blood-derived serine protease thrombin potentiates NMDA receptor response more than 2-fold through activation of PAR1 (37). Plasmin, another serine protease, similarly potentiates NMDA receptor response (38). Tissue-plasminogen activator (tPA), which catalyzes the conversion of the zymogen precursor plasminogen to plasmin and results in PAR1 activation, also interacts with and cleaves the ATD of the NR1 subunit of the NMDA receptor (39, 40). This raises the possibility that plasmin may also interact directly with the NMDA receptor subunits to modulate receptor response. We therefore investigated the ability of plasmin to cleave the NR2A NMDA receptor subunit. We found that nanomolar concentrations of plasmin can cleave within the ATD, a region that mediates tonic voltage-independent Zn2+ inhibition of NR2A-containing NMDA receptors. We hypothesized that plasmin cleavage reduces the Zn2+-mediated inhibition of NMDA receptors by removing the Zn2+ binding domain. In the present study, we have demonstrated that Zn2+ inhibition of agonist-evoked NMDA currents is decreased significantly by plasmin treatment in recombinant NR1/NR2A-transfected HEK 293 cells and cultured cortical neurons. These concentrations of plasmin may be pathophysiologically relevant in situations in which the blood-brain barrier is compromised, which could allow blood-derived plasmin to enter brain parenchyma at concentrations in excess of these that can cleave NR2A. Thus, ability of plasmin to potentiate NMDA function through the relief of the Zn2+ inhibition could exacerbate the harmful actions of NMDA receptor overactivation in pathological situations. In addition, if newly cleaved NR2AATD enters the bloodstream during ischemic injury, it could serve as a biomarker of central nervous system injury.  相似文献   

13.
Cold temperatures robustly activate a small cohort of somatosensory nerves, yet during a prolonged cold stimulus their activity will decrease, or adapt, over time. This process allows for the discrimination of subtle changes in temperature. At the molecular level, cold is detected by transient receptor potential melastatin 8 (TRPM8), a nonselective cation channel expressed on a subset of peripheral afferent fibers. We and others have reported that TRPM8 channels also adapt in a calcium-dependent manner when activated by the cooling compound menthol. Additionally, TRPM8 activity is sensitive to the phospholipid phosphoinositol 4,5-bisphosphate (PIP2), a substrate for the enzyme phospholipase C (PLC). These results suggest an adaptation model whereby TRPM8-mediated Ca2+ influx activates PLC, thereby decreasing PIP2 levels and resulting in reduced TRPM8 activity. Here we tested this model using pharmacological activation of PLC and by manipulating PIP2 levels independent of both PLC and Ca2+. PLC activation leads to adaptation-like reductions in cold- or menthol-evoked TRPM8 currents in both heterologous and native cells. Moreover, PLC-independent reductions in PIP2 had a similar effect on cold- and menthol-evoked currents. Mechanistically, either form of adaptation does not alter temperature sensitivity of TRPM8 but does lead to a change in channel gating. Our results show that adaptation is a shift in voltage dependence toward more positive potentials, reversing the trend toward negative potentials caused by agonist. These data suggest that PLC activity not only mediates adaptation to thermal stimuli, but likely underlies a more general mechanism that establishes the temperature sensitivity of somatosensory neurons.The detection of temperature is a fundamental task of the nervous system. Temperature-sensing sensory afferent neurons reside in either the trigeminal (TG)2 or dorsal root (DRG) sensory ganglia and project peripherally, terminating as free nerve endings that innervate areas of the head or trunk, respectively (1, 2). Subpopulations of these afferents respond to distinct sub-modalities of thermal stimuli, including noxious heat, innocuous cooling and warmth, and painfully cold temperatures. Each carries thermal information to the dorsal horn of the spinal cord, synapsing with neurons that project centrally (1, 3).The discovery of thermosensitive ion channels of the transient receptor potential (TRP) family demonstrated an underlying molecular mechanism for temperature detection (4). Cold temperature sensation is largely mediated by TRPM8, a nonselective cation channel expressed on a small subset of neurons (5, 6). TRPM8 is activated by cooling compounds, such as menthol, as well as cold temperatures below ∼28 °C, in vitro (7, 8). Recent reports on the behavioral phenotype of TRPM8-null mice suggest that this lone channel is required for the majority of cold sensing in vivo (5, 911). These and other data strongly implicate TRPM8 in not only the detection of both innocuous cool and some aspects of noxious cold but also injury-induced hypersensitivity to cold and, paradoxically, cooling-mediated analgesia (11, 12). Thus, understanding regulatory mechanisms that alter TRPM8 activity will provide keen insights into temperature sensation, nociception, and analgesia.One fundamental property of cold-sensitive neurons is an intrinsic ability to adapt to prolonged cold stimuli, a mechanism that is likely critical for discrimination of changing environmental conditions (13, 14). We and others have shown that cold-sensitive neurons adapt to cold and menthol over time in vitro (6, 15), a phenomenon also observed with recombinant TRPM8 channels activated by menthol (7). During sustained exposure to menthol, TRPM8 currents adapt in a manner that is dependent upon the presence of external calcium (7). Interestingly, cold- and menthol-evoked currents are highly sensitive to cellular manipulation. In heterologous cells, TRPM8 currents quickly decrease or run down upon membrane patch excision (16, 17). Moreover, in membrane patches excised from cold- and menthol-sensitive DRG neurons, cold thresholds for current activation exhibit a shift of ∼10 °C to colder temperatures in comparison with thresholds recorded in intact cells (18).Phosphatidylinositol 4,5-bisphosphate (PIP2) is a membrane phospholipid that accounts for ∼1% of all lipids in the inner leaflet of the plasma membrane and is known to regulate a variety of ion channels, including TRPM8 (16, 17). When applied to the cytoplasmic face of excised membrane patches containing TRPM8 channels, PIP2 can recover menthol-evoked currents to near pre-rundown levels (16, 17). PIP2 is proposed to interact with channels either through electrostatic interactions or by binding to target proteins at specific phosphoinositide-binding sites (19, 20). Membrane PIP2 levels are a product of enzymatic activity, such as phosphoinositide kinases that synthesize PIP2 from membrane precursors and phospholipase C (PLC) that hydrolyzes it, creating membrane-bound diacylglycerol (DAG) and cytosolic inositol trisphosphate (IP3), both of which function as second messengers. Of the three different PLC isotypes, PLCδ isoforms are modulated by increases in intracellular calcium (21).When taken in context with the sensitivity of TRPM8 currents to PIP2 levels, a model has been proposed whereby adaption is a result of channel-mediated Ca2+ influx activating one or more PLCδ isoforms (16, 17). The subsequent reductions in PIP2 levels then promote reduced or adapted TRPM8 currents. However, this hypothesis has not been conclusively shown in intact heterologous cells or in somatosensory neurons expressing TRPM8. Moreover, other alternative hypotheses for TRPM8 adaptation have been proposed, including Ca2+-dependent kinase activity mediated by protein kinase C (22, 23). Thus, the cellular and molecular mechanisms for Ca2+-mediated TRPM8 adaptation are unclear.Here we show, in both heterologous cells and native TRPM8-expressing neurons, that Ca2+-independent activation of PLC results in adapted TRPM8 currents. Moreover, PLC- and Ca2+-independent PIP2 depletion in heterologous cells produces similar effects on TRPM8 activity, again reducing both cold- and menthol-evoked currents. Mechanistically, we find that all such manipulations do not alter the temperature sensitivity of the channel but do lead to a shift in the voltage dependence of TRPM8 channel gating.  相似文献   

14.
Mitochondrial dysregulation is strongly implicated in Parkinson disease. Mutations in PTEN-induced kinase 1 (PINK1) are associated with familial parkinsonism and neuropsychiatric disorders. Although overexpressed PINK1 is neuroprotective, less is known about neuronal responses to loss of PINK1 function. We found that stable knockdown of PINK1 induced mitochondrial fragmentation and autophagy in SH-SY5Y cells, which was reversed by the reintroduction of an RNA interference (RNAi)-resistant plasmid for PINK1. Moreover, stable or transient overexpression of wild-type PINK1 increased mitochondrial interconnectivity and suppressed toxin-induced autophagy/mitophagy. Mitochondrial oxidant production played an essential role in triggering mitochondrial fragmentation and autophagy in PINK1 shRNA lines. Autophagy/mitophagy served a protective role in limiting cell death, and overexpressing Parkin further enhanced this protective mitophagic response. The dominant negative Drp1 mutant inhibited both fission and mitophagy in PINK1-deficient cells. Interestingly, RNAi knockdown of autophagy proteins Atg7 and LC3/Atg8 also decreased mitochondrial fragmentation without affecting oxidative stress, suggesting active involvement of autophagy in morphologic remodeling of mitochondria for clearance. To summarize, loss of PINK1 function elicits oxidative stress and mitochondrial turnover coordinated by the autophagic and fission/fusion machineries. Furthermore, PINK1 and Parkin may cooperate through different mechanisms to maintain mitochondrial homeostasis.Parkinson disease is an age-related neurodegenerative disease that affects ∼1% of the population worldwide. The causes of sporadic cases are unknown, although mitochondrial or oxidative toxins such as 1-methyl-4-phenylpyridinium, 6-hydroxydopamine (6-OHDA),3 and rotenone reproduce features of the disease in animal and cell culture models (1). Abnormalities in mitochondrial respiration and increased oxidative stress are observed in cells and tissues from parkinsonian patients (2, 3), which also exhibit increased mitochondrial autophagy (4). Furthermore, mutations in parkinsonian genes affect oxidative stress response pathways and mitochondrial homeostasis (5). Thus, disruption of mitochondrial homeostasis represents a major factor implicated in the pathogenesis of sporadic and inherited parkinsonian disorders (PD).The PARK6 locus involved in autosomal recessive and early-onset PD encodes for PTEN-induced kinase 1 (PINK1) (6, 7). PINK1 is a cytosolic and mitochondrially localized 581-amino acid serine/threonine kinase that possesses an N-terminal mitochondrial targeting sequence (6, 8). The primary sequence also includes a putative transmembrane domain important for orientation of the PINK1 domain (8), a conserved kinase domain homologous to calcium calmodulin kinases, and a C-terminal domain that regulates autophosphorylation activity (9, 10). Overexpression of wild-type PINK1, but not its PD-associated mutants, protects against several toxic insults in neuronal cells (6, 11, 12). Mitochondrial targeting is necessary for some (13) but not all of the neuroprotective effects of PINK1 (14), implicating involvement of cytoplasmic targets that modulate mitochondrial pathobiology (8). PINK1 catalytic activity is necessary for its neuroprotective role, because a kinase-deficient K219M substitution in the ATP binding pocket of PINK1 abrogates its ability to protect neurons (14). Although PINK1 mutations do not seem to impair mitochondrial targeting, PD-associated mutations differentially destabilize the protein, resulting in loss of neuroprotective activities (13, 15).Recent studies indicate that PINK1 and Parkin interact genetically (3, 16-18) to prevent oxidative stress (19, 20) and regulate mitochondrial morphology (21). Primary cells derived from PINK1 mutant patients exhibit mitochondrial fragmentation with disorganized cristae, recapitulated by RNA interference studies in HeLa cells (3).Mitochondria are degraded by macroautophagy, a process involving sequestration of cytoplasmic cargo into membranous autophagic vacuoles (AVs) for delivery to lysosomes (22, 23). Interestingly, mitochondrial fission accompanies autophagic neurodegeneration elicited by the PD neurotoxin 6-OHDA (24, 25). Moreover, mitochondrial fragmentation and increased autophagy are observed in neurodegenerative diseases including Alzheimer and Parkinson diseases (4, 26-28). Although inclusion of mitochondria in autophagosomes was once believed to be a random process, as observed during starvation, studies involving hypoxia, mitochondrial damage, apoptotic stimuli, or limiting amounts of aerobic substrates in facultative anaerobes support the concept of selective mitochondrial autophagy (mitophagy) (29, 30). In particular, mitochondrially localized kinases may play an important role in models involving oxidative mitochondrial injury (25, 31, 32).Autophagy is involved in the clearance of protein aggregates (33-35) and normal regulation of axonal-synaptic morphology (36). Chronic disruption of lysosomal function results in accumulation of subtly impaired mitochondria with decreased calcium buffering capacity (37), implicating an important role for autophagy in mitochondrial homeostasis (37, 38). Recently, Parkin, which complements the effects of PINK1 deficiency on mitochondrial morphology (3), was found to promote autophagy of depolarized mitochondria (39). Conversely, Beclin 1-independent autophagy/mitophagy contributes to cell death elicited by the PD toxins 1-methyl-4-phenylpyridinium and 6-OHDA (25, 28, 31, 32), causing neurite retraction in cells expressing a PD-linked mutation in leucine-rich repeat kinase 2 (40). Whereas properly regulated autophagy plays a homeostatic and neuroprotective role, excessive or incomplete autophagy creates a condition of “autophagic stress” that can contribute to neurodegeneration (28).As mitochondrial fragmentation (3) and increased mitochondrial autophagy (4) have been described in human cells or tissues of PD patients, we investigated whether or not the engineered loss of PINK1 function could recapitulate these observations in human neuronal cells (SH-SY5Y). Stable knockdown of endogenous PINK1 gave rise to mitochondrial fragmentation and increased autophagy and mitophagy, whereas stable or transient overexpression of PINK1 had the opposite effect. Autophagy/mitophagy was dependent upon increased mitochondrial oxidant production and activation of fission. The data indicate that PINK1 is important for the maintenance of mitochondrial networks, suggesting that coordinated regulation of mitochondrial dynamics and autophagy limits cell death associated with loss of PINK1 function.  相似文献   

15.
C-type lectin receptors expressed on the surface of dendritic cells and macrophages are able to bind glycoproteins of microbial pathogens via mannose, fucose, and N-acetylglucosamine. Langerin on Langerhans cells, dendritic cell-specific intercellular adhesion molecule 3-grabbing nonintegrin on dendritic cells, and mannose receptor (MR) on dendritic cells and macrophages bind the human immunodeficiency virus (HIV) envelope protein gp120 principally via high mannose oligosaccharides. These C-type lectin receptors can also oligomerize to facilitate enhanced ligand binding. This study examined the effect of oligomerization of MR on its ability to bind to mannan, monomeric gp120, native trimeric gp140, and HIV type 1 BaL. Mass spectrometry analysis of cross-linked MR showed homodimerization on the surface of primary monocyte-derived dendritic cells and macrophages. Both monomeric and dimeric MR were precipitated by mannan, but only the dimeric form was co-immunoprecipitated by gp120. These results were confirmed independently by flow cytometry analysis of soluble monomeric and trimeric HIV envelope and a cellular HIV virion capture assay. As expected, mannan bound to the carbohydrate recognition domains of MR dimers mostly in a calcium-dependent fashion. Unexpectedly, gp120-mediated binding of HIV to dimers on MR-transfected Rat-6 cells and macrophages was not calcium-dependent, was only partially blocked by mannan, and was also partially inhibited by N-acetylgalactosamine 4-sulfate. Thus gp120-mediated HIV binding occurs via the calcium-dependent, non-calcium-dependent carbohydrate recognition domains and the cysteine-rich domain at the C terminus of MR dimers, presenting a much broader target for potential inhibitors of gp120-MR binding.The mannose receptor (MR)2 is a C-type lectin receptor that is expressed on the surface of a variety of cells, including immature monocyte-derived dendritic cells (MDDC), dermal dendritic cells, macrophages, and hepatic endothelial cells. It is a multifunctional protein, involved in antigen recognition and internalization during the early stages of the innate immune response (1) as well as physiological clearance of the endogenous pituitary hormones lutropin and thyrotropin (2, 3). Recognition of foreign antigens occurs via mannose, fucose, and GlcNAc residues (4, 5), which are generally not found as terminal residues on mammalian glycoproteins but are highly abundant on surface proteins of pathogens such as the HIV-1 envelope gp120 (6, 7). Once bound, pathogens can be internalized by endocytosis or phagocytosis, where they are targeted to lysosomes for proteolytic degradation and presentation on major histocompatibility complex class II (8). In immature DCs, soluble recombinant HIV envelope proteins are processed by this pathway, initially binding to both dendritic cell-specific intracellular adhesion molecule 3 grabbing non-integrin (DC-SIGN) and MR and ultimately co-localizing with MR but not DC-SIGN in lysosomes (9). Furthermore, in immature DCs and to a greater extent mature DCs, a proportion of intact HIV-1 enters a unique vesicular compartment that co-localizes with tetraspanin proteins such as CD81 (10, 11). Recently, this compartment has been shown to be continuous with the plasma membrane (11) and does not represent a continuation of the endolysosomal network. Interestingly, this compartment can translocate virus from DCs to CD4 T cells, upon the formation of a virological synapse (1012). Although viral uptake can occur in DCs independent of HIV env (2), the efficiency of HIV binding and uptake is greatly enhanced by the presence of C-type lectin-env interactions. At least initial binding to DC-SIGN (and most likely also MR) is required for T cell trans-infection (13).Structurally, the extracellular domain of MR consists of an N-terminal cysteine-rich domain (Cys-RD), followed by a fibronectin type II domain and eight carbohydrate recognition domains (CRD) on a single polypeptide backbone (1). Of the eight CRDs, CRD 4–8 have been shown to be required for high affinity binding of ligands containing terminal mannose/fucose/GlcNAc residues, with CRD 4 having demonstrable monosaccharide binding in isolation (14). Binding and release of ligand within the low pH environment of the endolysosomal compartment are also Ca2+-dependent. Acid-induced removal of Ca2+ binding in CRD 4 and 5 was shown to cause a conformational rearrangement of the domain, resulting in a loss of carbohydrate binding activity (15). In contrast, binding of sulfated carbohydrates to the Cys-RD appears to be Ca2+-independent as no Ca2+-binding sites were observed in its crystal structure (2, 16).Oligomerization of CLRs such as DC-SIGN (17), Langerin (18), and mannose-binding protein (19) has been reported to be essential for binding of oligosaccharide-bearing ligands. Early studies on MR suggested that it exists solely as a monomeric molecule and that clustering of multiple CRDs within the single polypeptide backbone was necessary for high affinity binding of oligosaccharide moieties (20). However, more recent studies have shown that dimerization is possible in the presence of Ca2+ (21) and that an equilibrium may exist between monomeric and dimeric forms on the cell surface (22). It is currently unclear what effect dimerization has on ligand binding to the CRDs; however, there is evidence that dimerization of MR is required for high affinity binding of ligands bearing terminal N-acetylgalactosamine 4-sulfate (GalNAc-4-SO4) such as lutropin and thyrotropin (22) to the Cys-RD.To date, studies on the oligomerization and ligand binding activity of MR have used solubilized protein from cell lysates (20) or purified recombinant fragments (21). Because the membrane microenvironment can influence protein associations, soluble forms of MR may not necessarily be a true model of the quaternary structure and function of the native protein. Here, we used a well established method of cross-linking (23) on MDDCs, monocyte-derived macrophages (MDMs), and MR-transfected Rat-6 cells to preserve lateral protein-protein interactions between MR on the cell surface prior to solubilization. Mass spectrometry analysis of affinity-purified complexes showed they were homo-oligomers, and further resolution of the complex on a low percentage polyacrylamide gel by SDS-PAGE strongly indicates that they are dimers. Dimerization of MR was also found to be essential for binding mannan, monomeric gp120, native trimeric gp140, and HIV-1 viral particles. Persistence of monomeric gp120 and trimeric gp140 binding to dimeric MR in the presence of EGTA and various CRD and other inhibitors, however, suggested that gp120-mediated HIV-1 binding is not Ca2+-dependent and that at least binding probably occurs to both Ca2+-dependent and -independent CRDs and also the Cys-RD.  相似文献   

16.
Melanoma cells express the chemokine receptor CXCR4 that confers high invasiveness upon binding to its ligand CXCL12. Melanoma cells at initial stages of the disease show reduction or loss of E-cadherin expression, but recovery of its expression is frequently found at advanced phases. We overexpressed E-cadherin in the highly invasive BRO lung metastatic cell melanoma cell line to investigate whether it could influence CXCL12-promoted cell invasion. Overexpression of E-cadherin led to defective invasion of melanoma cells across Matrigel and type I collagen in response to CXCL12. A decrease in individual cell migration directionality toward the chemokine and reduced adhesion accounted for the impaired invasion. A p190RhoGAP-dependent inhibition of RhoA activation was responsible for the impairment in chemokine-stimulated E-cadherin melanoma transfectant invasion. Furthermore, we show that p190RhoGAP and p120ctn associated predominantly on the plasma membrane of cells overexpressing E-cadherin, and that E-cadherin-bound p120ctn contributed to RhoA inactivation by favoring p190RhoGAP-RhoA association. These results suggest that melanoma cells at advanced stages of the disease could have reduced metastatic potency in response to chemotactic stimuli compared with cells lacking E-cadherin, and the results indicate that p190RhoGAP is a central molecule controlling melanoma cell invasion.Cadherins are a family of Ca2+-dependent adhesion molecules that mediate cell-cell contacts and are expressed in most solid tissues providing a tight control of morphogenesis (1, 2). Classical cadherins, such as epithelial (E) cadherin, are found in adherens junctions, forming core protein complexes with β-catenin, α-catenin, and p120 catenin (p120ctn). Both β-catenin and p120ctn directly interact with E-cadherin, whereas α-catenin associates with the complex through its binding to β-catenin, providing a link with the actin cytoskeleton (1, 2). E-cadherin is frequently lost or down-regulated in many human tumors, coincident with morphological epithelial to mesenchymal transition and acquisition of invasiveness (3-6).Although melanoma only accounts for 5% of skin cancers, when metastasis starts, it is responsible for 80% of deaths from skin cancers (7). Melanocytes express E-cadherin (8-10), but melanoma cells at early radial growth phase show a large reduction in the expression of this cadherin, and surprisingly, expression has been reported to be partially recovered by vertical growth phase and metastatic melanoma cells (9, 11, 12).Trafficking of cancer cells from primary tumor sites to intravasation into blood circulation and later to extravasation to colonize distant organs requires tightly regulated directional cues and cell migration and invasion that are mediated by chemokines, growth factors, and adhesion molecules (13). Solid tumor cells express chemokine receptors that provide guidance of these cells to organs where their chemokine ligands are expressed, constituting a homing model resembling the one used by immune cells to exert their immune surveillance functions (14). Most solid cancer cells express CXCR4, a receptor for the chemokine CXCL12 (also called SDF-1), which is expressed in lungs, bone marrow, and liver (15). Expression of CXCR4 in human melanoma has been detected in the vertical growth phase and on regional lymph nodes, which correlated with poor prognosis and increased mortality (16, 17). Previous in vivo experiments have provided evidence supporting a crucial role for CXCR4 in the metastasis of melanoma cells (18).Rho GTPases control the dynamics of the actin cytoskeleton during cell migration (19, 20). The activity of Rho GTPases is tightly regulated by guanine-nucleotide exchange factors (GEFs),4 which stimulate exchange of bound GDP by GTP, and inhibited by GTPase-activating proteins (GAPs), which promote GTP hydrolysis (21, 22), whereas guanine nucleotide dissociation inhibitors (GDIs) appear to mediate blocking of spontaneous activation (23). Therefore, cell migration is finely regulated by the balance between GEF, GAP, and GDI activities on Rho GTPases. Involvement of Rho GTPases in cancer is well documented (reviewed in Ref. 24), providing control of both cell migration and growth. RhoA and RhoC are highly expressed in colon, breast, and lung carcinoma (25, 26), whereas overexpression of RhoC in melanoma leads to enhancement of cell metastasis (27). CXCL12 activates both RhoA and Rac1 in melanoma cells, and both GTPases play key roles during invasion toward this chemokine (28, 29).Given the importance of the CXCL12-CXCR4 axis in melanoma cell invasion and metastasis, in this study we have addressed the question of whether changes in E-cadherin expression on melanoma cells might affect cell invasiveness. We show here that overexpression of E-cadherin leads to impaired melanoma cell invasion to CXCL12, and we provide mechanistic characterization accounting for the decrease in invasion.  相似文献   

17.
Hepatocellular carcinoma (HCC) is one of the most common and aggressive human malignancies. Recombinant tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) is a promising anti-tumor agent. However, many HCC cells show resistance to TRAIL-induced apoptosis. In this study, we showed that bortezomib, a proteasome inhibitor, overcame TRAIL resistance in HCC cells, including Huh-7, Hep3B, and Sk-Hep1. The combination of bortezomib and TRAIL restored the sensitivity of HCC cells to TRAIL-induced apoptosis. Comparing the molecular change in HCC cells treated with these agents, we found that down-regulation of phospho-Akt (P-Akt) played a key role in mediating TRAIL sensitization of bortezomib. The first evidence was that bortezomib down-regulated P-Akt in a dose- and time-dependent manner in TRAIL-treated HCC cells. Second, LY294002, a PI3K inhibitor, also sensitized resistant HCC cells to TRAIL-induced apoptosis. Third, knocking down Akt1 by small interference RNA also enhanced TRAIL-induced apoptosis in Huh-7 cells. Finally, ectopic expression of mutant Akt (constitutive active) in HCC cells abolished TRAIL sensitization effect of bortezomib. Moreover, okadaic acid, a protein phosphatase 2A (PP2A) inhibitor, reversed down-regulation of P-Akt in bortezomib-treated cells, and PP2A knockdown by small interference RNA also reduced apoptosis induced by the combination of TRAIL and bortezomib, indicating that PP2A may be important in mediating the effect of bortezomib on TRAIL sensitization. Together, bortezomib overcame TRAIL resistance at clinically achievable concentrations in hepatocellular carcinoma cells, and this effect is mediated at least partly via inhibition of the PI3K/Akt pathway.Hepatocellular carcinoma (HCC)2 is currently the fifth most common solid tumor worldwide and the fourth leading cause of cancer-related death. To date, surgery is still the only curative treatment but is only feasible in a small portion of patients (1). Drug treatment is the major therapy for patients with advanced stage disease. Unfortunately, the response rate to traditional chemotherapy for HCC patients is unsatisfactory (1). Novel pharmacological therapy is urgently needed for patients with advanced HCC. In this regard, the approval of sorafenib might open a new era of molecularly targeted therapy in the treatment of HCC patients.Tumor necrosis factor (TNF)-related apoptosis-inducing ligand (TRAIL), a type II transmembrane protein and a member of the TNF family, is a promising anti-tumor agent under clinical investigation (2). TRAIL functions by engaging its receptors expressed on the surface of target cells. Five receptors specific for TRAIL have been identified, including DR4/TRAIL-R1, DR5/TRAIL-R2, DcR1, DcR2, and osteoprotegerin. Among TRAIL receptors, only DR4 and DR5 contain an effective death domain that is essential to formation of death-inducing signaling complex (DISC), a critical step for TRAIL-induced apoptosis. Notably, the trimerization of the death domains recruits an adaptor molecule, Fas-associated protein with death domain (FADD), which subsequently recruits and activates caspase-8. In type I cells, activation of caspase-8 is sufficient to activate caspase-3 to induce apoptosis; however, in another type of cells (type II), the intrinsic mitochondrial pathway is essential for apoptosis characterized by cleavage of Bid and release of cytochrome c from mitochondria, which subsequently activates caspase-9 and caspase-3 (3).Although TRAIL induces apoptosis in malignant cells but sparing normal cells, some tumor cells are resistant to TRAIL-induced apoptosis. Mechanisms responsible for the resistance include receptors and intracellular resistance. Although the cell surface expression of DR4 or DR5 is absolutely required for TRAIL-induced apoptosis, tumor cells expressing these death receptors are not always sensitive to TRAIL due to intracellular mechanisms. For example, the cellular FLICE-inhibitory protein (c-FLIP), a homologue to caspase-8 but without protease activity, has been linked to TRAIL resistance in several studies (4, 5). In addition, inactivation of Bax, a proapoptotic Bcl-2 family protein, resulted in resistance to TRAIL in MMR-deficient tumors (6, 7), and reintroduction of Bax into Bax-deficient cells restored TRAIL sensitivity (8), indicating that the Bcl-2 family plays a critical role in intracellular mechanisms for resistance of TRAIL.Bortezomib, a proteasome inhibitor approved clinically for multiple myeloma and mantle cell lymphoma, has been investigated intensively for many types of cancer (9). Accumulating studies indicate that the combination of bortezomib and TRAIL overcomes the resistance to TRAIL in various types of cancer, including acute myeloid leukemia (4), lymphoma (1013), prostate (1417), colon (15, 18, 19), bladder (14, 16), renal cell carcinoma (20), thyroid (21), ovary (22), non-small cell lung (23, 24), sarcoma (25), and HCC (26, 27). Molecular targets responsible for the sensitizing effect of bortezomib on TRAIL-induced cell death include DR4 (14, 27), DR5 (14, 20, 2223, 28), c-FLIP (4, 11, 2123, 29), NF-κB (12, 24, 30), p21 (16, 21, 25), and p27 (25). In addition, Bcl-2 family also plays a role in the combinational effect of bortezomib and TRAIL, including Bcl-2 (10, 21), Bax (13, 22), Bak (27), Bcl-xL (21), Bik (18), and Bim (15).Recently, we have reported that Akt signaling is a major molecular determinant in bortezomib-induced apoptosis in HCC cells (31). In this study, we demonstrated that bortezomib overcame TRAIL resistance in HCC cells through inhibition of the PI3K/Akt pathway.  相似文献   

18.
19.
Rapid protein kinase D (PKD) activation and phosphorylation via protein kinase C (PKC) have been extensively documented in many cell types cells stimulated by multiple stimuli. In contrast, little is known about the role and mechanism(s) of a recently identified sustained phase of PKD activation in response to G protein-coupled receptor agonists. To elucidate the role of biphasic PKD activation, we used Swiss 3T3 cells because PKD expression in these cells potently enhanced duration of ERK activation and DNA synthesis in response to Gq-coupled receptor agonists. Cell treatment with the preferential PKC inhibitors GF109203X or Gö6983 profoundly inhibited PKD activation induced by bombesin stimulation for <15 min but did not prevent PKD catalytic activation induced by bombesin stimulation for longer times (>60 min). The existence of sequential PKC-dependent and PKC-independent PKD activation was demonstrated in 3T3 cells stimulated with various concentrations of bombesin (0.3–10 nm) or with vasopressin, a different Gq-coupled receptor agonist. To gain insight into the mechanisms involved, we determined the phosphorylation state of the activation loop residues Ser744 and Ser748. Transphosphorylation targeted Ser744, whereas autophosphorylation was the predominant mechanism for Ser748 in cells stimulated with Gq-coupled receptor agonists. We next determined which phase of PKD activation is responsible for promoting enhanced ERK activation and DNA synthesis in response to Gq-coupled receptor agonists. We show, for the first time, that the PKC-independent phase of PKD activation mediates prolonged ERK signaling and progression to DNA synthesis in response to bombesin or vasopressin through a pathway that requires epidermal growth factor receptor-tyrosine kinase activity. Thus, our results identify a novel mechanism of Gq-coupled receptor-induced mitogenesis mediated by sustained PKD activation through a PKC-independent pathway.The understanding of the mechanisms that control cell proliferation requires the identification of the molecular pathways that govern the transition of quiescent cells into the S phase of the cell cycle. In this context the activation and phosphorylation of protein kinase D (PKD),4 the founding member of a new protein kinase family within the Ca2+/calmodulin-dependent protein kinase (CAMK) group and separate from the previously identified PKCs (for review, see Ref. 1), are attracting intense attention. In unstimulated cells, PKD is in a state of low catalytic (kinase) activity maintained by autoinhibition mediated by the N-terminal domain, a region containing a repeat of cysteinerich zinc finger-like motifs and a pleckstrin homology (PH) domain (14). Physiological activation of PKD within cells occurs via a phosphorylation-dependent mechanism first identified in our laboratory (57). In response to cellular stimuli (1), including phorbol esters, growth factors (e.g. PDGF), and G protein-coupled receptor (GPCR) agonists (6, 816) that signal through Gq, G12, Gi, and Rho (11, 1519), PKD is converted into a form with high catalytic activity, as shown by in vitro kinase assays performed in the absence of lipid co-activators (5, 20).During these studies multiple lines of evidence indicated that PKC activity is necessary for rapid PKD activation within intact cells. For example, rapid PKD activation was selectively and potently blocked by cell treatment with preferential PKC inhibitors (e.g. GF109203X or Gö6983) that do not directly inhibit PKD catalytic activity (5, 20), implying that PKD activation in intact cells is mediated directly or indirectly through PKCs. Many reports demonstrated the operation of a rapid PKC/PKD signaling cascade induced by multiple GPCR agonists and other receptor ligands in a range of cell types (for review, see Ref. 1). Our previous studies identified Ser744 and Ser748 in the PKD activation loop (also referred as activation segment or T-loop) as phosphorylation sites critical for PKC-mediated PKD activation (1, 4, 7, 17, 21). Collectively, these findings demonstrated the existence of a rapidly activated PKC-PKD protein kinase cascade(s). In a recent study we found that the rapid PKC-dependent PKD activation was followed by a late, PKC-independent phase of catalytic activation and phosphorylation induced by stimulation of the bombesin Gq-coupled receptor ectopically expressed in COS-7 cells (22). This study raised the possibility that PKD mediates rapid biological responses downstream of PKCs, whereas, in striking contrast, PKD could mediate long term responses through PKC-independent pathways. Despite its potential importance for defining the role of PKC and PKD in signal transduction, this hypothesis has not been tested in any cell type.Accumulating evidence demonstrates that PKD plays an important role in several cellular processes and activities, including signal transduction (14, 2325), chromatin organization (26), Golgi function (27, 28), gene expression (2931), immune regulation (26), and cell survival, adhesion, motility, differentiation, DNA synthesis, and proliferation (for review, see Ref. 1). In Swiss 3T3 fibroblasts, a cell line used extensively as a model system to elucidate mechanisms of mitogenic signaling (3234), PKD expression potently enhances ERK activation, DNA synthesis, and cell proliferation induced by Gq-coupled receptor agonists (8, 14). Here, we used this model system to elucidate the role and mechanism(s) of biphasic PKD activation. First, we show that the Gq-coupled receptor agonists bombesin and vasopressin, in contrast to phorbol esters, specifically induce PKD activation through early PKC-dependent and late PKC-independent mechanisms in Swiss 3T3 cells. Subsequently, we demonstrate for the first time that the PKC-independent phase of PKD activation is responsible for promoting ERK signaling and progression to DNA synthesis through an epidermal growth factor receptor (EGFR)-dependent pathway. Thus, our results identify a novel mechanism of Gq-coupled receptor-induced mitogenesis mediated by sustained PKD activation through a PKC-independent pathway.  相似文献   

20.
Galectins have been implicated in T cell homeostasis playing complementary pro-apoptotic roles. Here we show that galectin-8 (Gal-8) is a potent pro-apoptotic agent in Jurkat T cells inducing a complex phospholipase D/phosphatidic acid signaling pathway that has not been reported for any galectin before. Gal-8 increases phosphatidic signaling, which enhances the activity of both ERK1/2 and type 4 phosphodiesterases (PDE4), with a subsequent decrease in basal protein kinase A activity. Strikingly, rolipram inhibition of PDE4 decreases ERK1/2 activity. Thus Gal-8-induced PDE4 activation releases a negative influence of cAMP/protein kinase A on ERK1/2. The resulting strong ERK1/2 activation leads to expression of the death factor Fas ligand and caspase-mediated apoptosis. Several conditions that decrease ERK1/2 activity also decrease apoptosis, such as anti-Fas ligand blocking antibodies. In addition, experiments with freshly isolated human peripheral blood mononuclear cells, previously stimulated with anti-CD3 and anti-CD28, show that Gal-8 is pro-apoptotic on activated T cells, most likely on a subpopulation of them. Anti-Gal-8 autoantibodies from patients with systemic lupus erythematosus block the apoptotic effect of Gal-8. These results implicate Gal-8 as a novel T cell suppressive factor, which can be counterbalanced by function-blocking autoantibodies in autoimmunity.Glycan-binding proteins of the galectin family have been increasingly studied as regulators of the immune response and potential therapeutic agents for autoimmune disorders (1). To date, 15 galectins have been identified and classified according with the structural organization of their distinctive monomeric or dimeric carbohydrate recognition domain for β-galactosides (2, 3). Galectins are secreted by unconventional mechanisms and once outside the cells bind to and cross-link multiple glycoconjugates both at the cell surface and at the extracellular matrix, modulating processes as diverse as cell adhesion, migration, proliferation, differentiation, and apoptosis (410). Several galectins have been involved in T cell homeostasis because of their capability to kill thymocytes, activated T cells, and T cell lines (1116). Pro-apoptotic galectins might contribute to shape the T cell repertoire in the thymus by negative selection, restrict the immune response by eliminating activated T cells at the periphery (1), and help cancer cells to escape the immune system by eliminating cancer-infiltrating T cells (17). They have also a promising therapeutic potential to eliminate abnormally activated T cells and inflammatory cells (1). Studies on the mostly explored galectins, Gal-1, -3, and -9 (14, 15, 1820), as well as in Gal-2 (13), suggest immunosuppressive complementary roles inducing different pathways to apoptosis. Galectin-8 (Gal-8)4 is one of the most widely expressed galectins in human tissues (21, 22) and cancerous cells (23, 24). Depending on the cell context and mode of presentation, either as soluble stimulus or extracellular matrix, Gal-8 can promote cell adhesion, spreading, growth, and apoptosis (6, 7, 9, 10, 22, 25). Its role has been mostly studied in relation to tumor malignancy (23, 24). However, there is some evidence regarding a role for Gal-8 in T cell homeostasis and autoimmune or inflammatory disorders. For instance, the intrathymic expression and pro-apoptotic effect of Gal-8 upon CD4highCD8high thymocytes suggest a role for Gal-8 in shaping the T cell repertoire (16). Gal-8 could also modulate the inflammatory function of neutrophils (26), Moreover Gal-8-blocking agents have been detected in chronic autoimmune disorders (10, 27, 28). In rheumatoid arthritis, Gal-8 has an anti-inflammatory action, promoting apoptosis of synovial fluid cells, but can be counteracted by a specific rheumatoid version of CD44 (CD44vRA) (27). In systemic lupus erythematosus (SLE), a prototypic autoimmune disease, we recently described function-blocking autoantibodies against Gal-8 (10, 28). Thus it is important to define the role of Gal-8 and the influence of anti-Gal-8 autoantibodies in immune cells.In Jurkat T cells, we previously reported that Gal-8 interacts with specific integrins, such as α1β1, α3β1, and α5β1 but not α4β1, and as a matrix protein promotes cell adhesion and asymmetric spreading through activation of the extracellular signal-regulated kinases 1 and 2 (ERK1/2) (10). These early effects occur within 5–30 min. However, ERK1/2 signaling supports long term processes such as T cell survival or death, depending on the moment of the immune response. During T cell activation, ERK1/2 contributes to enhance the expression of interleukin-2 (IL-2) required for T cell clonal expansion (29). It also supports T cell survival against pro-apoptotic Fas ligand (FasL) produced by themselves and by other previously activated T cells (30, 31). Later on, ERK1/2 is required for activation-induced cell death, which controls the extension of the immune response by eliminating recently activated and restimulated T cells (32, 33). In activation-induced cell death, ERK1/2 signaling contributes to enhance the expression of FasL and its receptor Fas/CD95 (32, 33), which constitute a preponderant pro-apoptotic system in T cells (34). Here, we ask whether Gal-8 is able to modulate the intensity of ERK1/2 signaling enough to participate in long term processes involved in T cell homeostasis.The functional integration of ERK1/2 and PKA signaling (35) deserves special attention. cAMP/PKA signaling plays an immunosuppressive role in T cells (36) and is altered in SLE (37). Phosphodiesterases (PDEs) that degrade cAMP release the immunosuppressive action of cAMP/PKA during T cell activation (38, 39). PKA has been described to control the activity of ERK1/2 either positively or negatively in different cells and processes (35). A little explored integration among ERK1/2 and PKA occurs via phosphatidic acid (PA) and PDE signaling. Several stimuli activate phospholipase D (PLD) that hydrolyzes phosphatidylcholine into PA and choline. Such PLD-generated PA plays roles in signaling interacting with a variety of targeting proteins that bear PA-binding domains (40). In this way PA recruits Raf-1 to the plasma membrane (41). It is also converted by phosphatidic acid phosphohydrolase (PAP) activity into diacylglycerol (DAG), which among other functions, recruits and activates the GTPase Ras (42). Both Ras and Raf-1 are upstream elements of the ERK1/2 activation pathway (43). In addition, PA binds to and activates PDEs of the type 4 subfamily (PDE4s) leading to decreased cAMP levels and PKA down-regulation (44). The regulation and role of PA-mediated control of ERK1/2 and PKA remain relatively unknown in T cell homeostasis, because it is also unknown whether galectins stimulate the PLD/PA pathway.Here we found that Gal-8 induces apoptosis in Jurkat T cells by triggering cross-talk between PKA and ERK1/2 pathways mediated by PLD-generated PA. Our results for the first time show that a galectin increases the PA levels, down-regulates the cAMP/PKA system by enhancing rolipram-sensitive PDE activity, and induces an ERK1/2-dependent expression of the pro-apoptotic factor FasL. The enhanced PDE activity induced by Gal-8 is required for the activation of ERK1/2 that finally leads to apoptosis. Gal-8 also induces apoptosis in human peripheral blood mononuclear cells (PBMC), especially after activating T cells with anti-CD3/CD28. Therefore, Gal-8 shares with other galectins the property of killing activated T cells contributing to the T cell homeostasis. The pathway involves a particularly integrated signaling context, engaging PLD/PA, cAMP/PKA, and ERK1/2, which so far has not been reported for galectins. The pro-apoptotic function of Gal-8 also seems to be unique in its susceptibility to inhibition by anti-Gal-8 autoantibodies.  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