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

2.
3.
Relationships among biochemical signaling processes involved in Ca2+/calmodulin (CaM)-dependent phosphorylation of smooth muscle myosin regulatory light chain (RLC) by myosin light chain kinase (MLCK) were determined. A genetically-encoded biosensor MLCK for measuring Ca2+-dependent CaM binding and activation was expressed in smooth muscles of transgenic mice. We performed real-time evaluations of the relationships among [Ca2+]i, MLCK activation, and contraction in urinary bladder smooth muscle strips neurally stimulated for 3 s. Latencies for the onset of [Ca2+]i and kinase activation were 55 ± 8 and 65 ± 6 ms, respectively. Both increased with RLC phosphorylation at 100 ms, whereas force latency was 109 ± 3 ms. [Ca2+]i, kinase activation, and RLC phosphorylation responses were maximal by 1.2 s, whereas force increased more slowly to a maximal value at 3 s. A delayed temporal response between RLC phosphorylation and force is probably due to mechanical effects associated with elastic elements in the tissue. MLCK activation partially declined at 3 s of stimulation with no change in [Ca2+]i and also declined more rapidly than [Ca2+]i during relaxation. The apparent desensitization of MLCK to Ca2+ activation appears to be due to phosphorylation in its calmodulin binding segment. Phosphorylation of two myosin light chain phosphatase regulatory proteins (MYPT1 and CPI-17) or a protein implicated in strengthening membrane adhesion complexes for force transmission (paxillin) did not change during force development. Thus, neural stimulation leads to rapid increases in [Ca2+]i, MLCK activation, and RLC phosphorylation in phasic smooth muscle, showing a tightly coupled Ca2+ signaling complex as an elementary mechanism initiating contraction.Increases in [Ca2+]i3 in smooth muscle cells lead to Ca2+/CaM-dependent MLCK activation and RLC phosphorylation. Phosphorylation of RLC increases actin-activated myosin MgATPase activity leading to myosin cross-bridge cycling with force development (13).The activation of smooth muscle contraction may be affected by multiple cellular processes. Previous investigations show that free Ca2+/CaM is limiting for kinase activation despite the abundance of total CaM (46). The extent of RLC phosphorylation is balanced by the actions of MLCK and myosin light chain phosphatase, which is composed of three distinct protein subunits (7). The myosin phosphatase targeting subunit, MYPT1, in smooth muscle binds to myosin filaments, thus targeting the 37-kDa catalytic subunit (type 1 serine/threonine phosphatase, PP1c) to phosphorylated RLC. RLC phosphorylation and muscle force may be regulated by additional signaling pathways involving phosphorylation of RLC by Ca2+-independent kinase(s) and inhibition of myosin light chain phosphatase, processes that increase the contraction response at fixed [Ca2+]i (Ca2+-sensitization) (814). Many studies indicate that agonist-mediated Ca2+-sensitization most often reflects decreased myosin light chain phosphatase activity involving two major pathways including MYPT1 phosphorylation by a Rho kinase pathway and phosphorylation of CPI-17 by PKC (8, 1416). Additionally, phosphorylation of MLCK in its calmodulin-binding sequence by a Ca2+/calmodulin-dependent kinase pathway has been implicated in Ca2+ desensitization of RLC phosphorylation (1719). How these signaling pathways intersect the responses of the primary Ca2+/CaM pathway during physiological neural stimulation is not known.There is also evidence that smooth muscle contraction requires the polymerization of submembranous cytoskeletal actin filaments to strengthen membrane adhesion complexes involved in transmitting force between actin-myosin filaments and external force-transmitting structures (2023). In tracheal smooth muscle, paxillin at membrane adhesions undergoes tyrosine phosphorylation in response to contractile stimulation by an agonist, and this phosphorylation increases concurrently with force development in response to agonist. Expression of nonphosphorylatable paxillin mutants in tracheal muscle suppresses acetylcholine-induced tyrosine phosphorylation of paxillin, tension development, and actin polymerization without affecting RLC phosphorylation (24, 25). Thus, paxillin phosphorylation may play an important role in tension development in smooth muscle independently of RLC phosphorylation and cross-bridge cycling.Specific models relating signaling mechanisms in the smooth muscle cell to contraction dynamics are limited when cells in tissues are stimulated slowly and asynchronously by agonist diffusing into the preparation. Field stimulation leading to the rapid release of neurotransmitters from nerves embedded in the tissue avoids these problems associated with agonist diffusion (26, 27). In urinary bladder smooth muscle, phasic contractions are brought about by the parasympathetic nervous system. Upon activation, parasympathetic nerve varicosities release the two neurotransmitters, acetylcholine and ATP, that bind to muscarinic and purinergic receptors, respectively. They cause smooth muscle contraction by inducing Ca2+ transients as elementary signals in the process of nerve-smooth muscle communication (2830). We recently reported the development of a genetically encoded, CaM-sensor for activation of MLCK. The CaM-sensor MLCK contains short smooth muscle MLCK fused to two fluorophores, enhanced cyan fluorescent protein and enhanced yellow fluorescent protein, linked by the MLCK calmodulin binding sequence (6, 14, 31). Upon dimerization, there is significant FRET from the donor enhanced cyan fluorescent protein to the acceptor enhanced yellow fluorescent protein. Ca2+/CaM binding dissociates the dimer resulting in a decrease in FRET intensity coincident with activation of kinase activity (31). Thus, CaM-sensor MLCK is capable of directly monitoring Ca2+/CaM binding and activation of the kinase in smooth muscle tissues where it is expressed specifically in smooth muscle cells of transgenic mice. We therefore combined neural stimulation with real-time measurements of [Ca2+]i, MLCK activation, and force development in smooth muscle tissue from these mice. Additionally, RLC phosphorylation was measured precisely at specific times following neural stimulation in tissues frozen by a rapid-release electronic freezing device (26, 27). Results from these studies reveal that physiological stimulation of smooth muscle cells by neurotransmitter release leads to rapid increases in [Ca2+]i, MLCK activation, and RLC phosphorylation at similar rates without the apparent activities of Ca2+-independent kinases, inhibition of myosin light chain phosphatase, or paxillin phosphorylation. Thus, the elemental processes for phasic smooth muscle contraction are represented by this tightly coupled Ca2+ signaling complex.  相似文献   

4.
5.
The ubiquitous Ca2+-sensing protein calmodulin (CaM) fulfills its numerous signaling functions through a wide range of modular binding and activation mechanisms. By activating adenylyl cyclases (ACs) 1 and 8, Ca2+ acting via calmodulin impacts on the signaling of the other major cellular second messenger cAMP. In possessing two CaM-binding domains, a 1-5-8-14 motif at the N terminus and an IQ-like motif (IQlm) at the C terminus, AC8 offers particularly sophisticated regulatory possibilities. The IQlm has remained unexplored beyond the suggestion that it bound CaM, and the larger C2b region of which it is part was involved in the relief of autoinhibition of AC8. Here we attempt to distinguish the function of individual residues of the IQlm. From a complementary approach of in vitro and cell population AC activity assays, as well as CaM binding, we propose that the IQlm alone, and not the majority of the C2b, imparts CaM binding and autoinhibitory functions. Moreover, this duality of function is spatially separated and depends on amino acid side-chain character. Accordingly, residues critical for CaM binding are positively charged and clustered toward the C terminus, and those essential for the maintenance of autoinhibition are hydrophobic and more N-terminal. Secondary structure prediction of the IQlm supports this separation, with an ideally placed break in the α-helical nature of the sequence. We additionally find that the N and C termini of AC8 interact, which is an association specifically abrogated by fully Ca2+-bound, but not Ca2+-free, CaM. These data support a sophisticated activation mechanism of AC8 by CaM, in which the duality of the IQlm function is critical.The divalent calcium ion, Ca2+, plays a key role in modulating cellular processes as diverse as fertilization and apoptosis (1, 2). Ca2+ concentrations inside the cell are held ∼100 nm, despite a perpetually higher level of 1–2 mm in the extracellular medium. This steep gradient across the plasma membrane allows for a large influx when Ca2+ channels open, with subsequent signaling events that often rely on transduction via Ca2+-binding proteins (3). The archetypal Ca2+ sensor is calmodulin (CaM),2 a small, acidic protein so strictly conserved that all vertebrate CaM genes encode an identical 148-residue sequence (4). Multifunctional in its downstream effects, CaM can bind to at least 300 target proteins with novel partners continuing to be discovered (5). The list of effectors includes two isoforms of the adenylyl cyclase (AC) superfamily, AC1 and AC8, which comprise the Ca2+-stimulable subset of the nine particulate ACs (6). In intact cells, AC8 exhibits a predilection for store-operated, or capacitative, Ca2+ entry (CCE) (7, 8). This mode of Ca2+ entry is triggered by the emptying of endo/sarcoplasmic reticular stores by physiological or pharmacological stimuli (9, 10). Although the mechanism of the regulation of AC8 in nonexcitable cells by CCE (or voltage-gated Ca2+ entry in neurons) can be viewed to rely on facets of cellular compartmentalization (1113), the detailed molecular mechanism whereby Ca2+ stimulates the enzyme is unclear. Consequently, the present investigation addresses the molecular mechanism whereby Ca2+, acting via calmodulin, stimulates AC8.The broad features of CaM that hold the key to its multiple regulatory strategies are understood. CaM is organized into two homologous globular domains (or lobes) united by a short linker segment (14). Both the N-terminal (N-lobe) and the C-terminal lobe (C-lobe) include two EF-hands (specialized Ca2+-coordinating helix-loop-helix motifs (15)), endowing CaM with four Ca2+-binding sites. Ca2+ binding to either lobe of CaM induces a structural reconfiguration determined by the two helices of each EF-hand separating from near-antiparallel to perpendicular arrays (14, 16). This exposes hydrophobic trenches in the C- and N-lobes sequentially (because the former has the highest affinity for Ca2+), which are notably lined with a disproportionate number of flexible methionine side chains, ready to accommodate a remarkable array of unrelated sequences. Initially, the mode of Ca2+-loaded CaM association with targets was considered to be uniform; the N- and C-lobes collapsed around the target peptides of e.g. smooth muscle myosin light chain kinase (17), skeletal muscle myosin light chain kinase (18), and CaMKIIα (19) with the N-lobe favoring the C-terminal target sequence and vice versa. However, CaM has since proven to be more versatile and unpredictable in how it associates with and regulates effectors, with reports of N-lobe-N-terminal/C-lobe-C-terminal interaction (20), target dimerization promoted by CaM (21), CaM binding to fatty acyl modifications (22), and other deviations from the early model.Nevertheless, three main forms of CaM regulation have been proposed as follows: relief of autoinhibition; active site remodeling; and dimerization of target domains (23). Whether the precise mechanism of CaM binding and subsequent regulation of AC8 falls into these categories of CaM regulation is not resolved. A previous study (24) established that AC8 possesses two calmodulin-binding domains (CaMBDs). CaM recognition sequences generally show little homology, although classifications based on relative positions of key hydrophobic residues have been usefully applied (4). The N-terminal CaMBD of AC8 conforms to a “1-5-8-14 motif” having large hydrophobic side chains from Trp, Val, or Ile residues at these spatially conserved sites. The C-terminal CaMBD contains an IQ-like motif (IQlm), in accordance with a signature (IVL)QXXXR(K) arrangement, to which CaM binds directly (25). By truncating one terminus or both termini, Gu and Cooper (24) asserted that the N terminus contributed little to direct CaM regulation of AC8, whereas the C terminus was critical for the maintenance of an auto-inhibited state, which was relieved upon binding of CaM. Thus, functional elements of both autoinhibition and pre-association, imparted by noncontiguous CaMBDs, have been suggested, thereby excluding AC8 from a simple model of CaM binding to an autoinhibitory domain leading to activation, as is sometimes observed (23).Recently, the proposal that the two CaMBDs play separate roles in AC8 activation was reinforced (26). This study suggested that CaM tethering by the N-terminal 1-5-8-14 motif provides the catalytically relevant C-terminal CaMBD with privileged access to CaM, thereby circumventing the need for AC8 to compete for CaM, whose free concentration in the cell is far exceeded by that of its targets (27, 28). 1–14 motifs are more generally employed in relieving autoinhibitory influences (23), so the use by AC8 of a 1-5-8-14 motif as a CaM-tethering site is unusual. The proposal that CaM pre-associates here was based on limited mutagenesis of residues within the 1-5-8-14 motif of AC8 and the use of CaM mutants whose Ca2+ binding capability was abolished completely or limited to discrete lobes. In contrast to the 1-5-8-14 motif at the N terminus, very little is known of the IQlm at the C terminus, in terms of the roles of the individual residues in CaM binding or autoinhibition, or even on the interplay between CaMBDs at the N and C termini of AC8.Against this background, the present study sought to assess the contribution of IQlm residues to the function of AC8, focusing on consequences of key mutations on CaM-binding efficiency, regulation by Ca2+/CaM, and maintenance of the autoinhibited state. Through this series of experiments, the level of coordination between the IQ-like and 1-5-8-14 motifs became evident. The provision of a pre-associated CaM molecule was found to allow for potentially deleterious mutations of the IQlm to be tolerated. Within this latter motif, Leu1196, Val1197, and Leu1200 are residues essential to autoinhibition, whereas the main responsibility of binding CaM directly at the C terminus lies with Arg1202 and Arg1204. In this regard, the AC8 IQlm spatially separates the two functions of CaM binding and maintenance of autoinhibition, a separation that is supported by a predicted break in the helicity of the IQlm region. Thus, a new variation is revealed in the manner by which a target exploits the CaM device into a sophisticated activation mechanism.  相似文献   

6.
Ca2+/calmodulin-dependent protein kinase II (αCaMKII) is thought to exert its role in memory formation by autonomous Ca2+-independent persistent activity conferred by Thr286 autophosphorylation, allowing the enzyme to remain active even when intracellular [Ca2+] has returned to resting levels. Ca2+ sequestration-induced inhibition, caused by a burst of Thr305/306 autophosphorylation via calmodulin (CaM) dissociation from the Thr305/306 sites, is in conflict with this view. The processes of CaM binding, autophosphorylation, and inactivation are dissected to resolve this conflict. Upon Ca2+ withdrawal, CaM sequential domain dissociation is observed, starting with the rapid release of the first (presumed N-terminal) CaM lobe, thought to be bound at the Thr305/306 sites. The time courses of Thr305/306 autophosphorylation and inactivation, however, correlate with the slow dissociation of the second (presumed C-terminal) CaM lobe. Exposure of the Thr305/306 sites is thus not sufficient for their autophosphorylation. Moreover, Thr305/306 autophosphorylation and autoinactivation are shown to occur in the continuous presence of Ca2+ and bound Ca2+/CaM by time courses similar to those seen following Ca2+ sequestration. Our investigation of the activity and mechanisms of phospho-Thr286-αCaMKII thus shows time-dependent autoinactivation, irrespective of the continued presence of Ca2+ and CaM, allowing a very short, if any, time window for Ca2+/CaM-free phospho-Thr286-αCaMKII activity. Physiologically, the time-dependent autoinactivation mechanisms of phospho-Thr286-αCaMKII (t½ of ∼50 s at 37 °C) suggest a transient kinase activity of ∼1 min duration in the induction of long term potentiation and thus memory formation.Ca2+/calmodulin-dependent protein kinase II (αCaMKII)2 is essential in hippocampal learning and N-methyl-d-aspartate receptor-dependent synaptic plasticity, causing long term potentiation (1, 2). The exact mechanisms of αCaMKII in memory functions have not yet been identified.αCaMKII is a broad specificity Ser/Thr protein kinase, which catalyzes the phosphorylation of over 100 protein and peptide substrates in vitro (3). Uniquely, the CaMKII family possesses two distinct kinase mechanisms. The first mechanism is a “canonical” intrasubunit phosphorylation, commonly found in monomeric kinases, in which the phosphorylatable residue of the substrate bound to the helical subdomain of the catalytic domain at the active site is lined up with the terminal phosphate of ATP (4). Although there is a large number of potential “canonical” substrates for αCaMKII at the synapse (5), so far AMPA receptors have been shown to be possible physiological substrates of αCaMKII (6). For the purpose of this study, syntide 2, a commonly used peptide substrate derived from phosphorylation site 2 of glycogen synthase (7), was chosen.The second mechanism, intersubunit autophosphorylation, takes advantage of the oligomeric organization of CaMKII (8). The most important autophosphorylation site in the α isoform is Thr286, which resides in the vicinity of the autoinhibitory domain (9). Peptide substrates with homologous sequences to this region have been reported to be phosphorylated by αCaMKII. This, however, occurs with a low Vmax, and these substrates show properties of a non-competitive inhibitor with respect to phosphorylation of “canonical” substrates (10) and of Thr286 autophosphorylation itself (11). Examples of such substrates include autocamtide, a peptide substrate derived from the autoinhibitory region (12) and the NR2B subunit of the N-methyl-d-aspartate receptor, which has been identified as a potential physiological target of phospho-Thr286-αCaMKII at the postsynaptic membrane (13). The possible physiological significance of NR2B phosphorylation is not yet known. There is evidence to suggest that Thr286 autophosphorylation is required to achieve full activity of the enzyme, since the unphosphorylatable T286A mutant enzyme has much diminished activity compared with wild type enzyme (14, 15).Thr286 autophosphorylation causes CaM “trapping,” a >104-fold increase in the affinity of αCaMKII for Ca2+/CaM (1618). At the same time, Thr286 autophosphorylation is also attributed to confer Ca2+- and CaM-independent persistent “autonomous” kinase activity to αCaMKII. However, due to the extremely high affinity of phospho-Thr286-αCaMKII for Ca2+/CaM, [Ca2+] of <10 nm is required to achieve full dissociation of Ca2+/CaM, since CaM trapping occurs by virtue of Ca2+ trapping (19). Partial activity measured upon partial Ca2+ withdrawal therefore may not always reflect Ca2+/CaM-free enzyme (9). Furthermore, the physiological resting [Ca2+] range is 50–100 nm; therefore, phospho-Thr286-αCaMKII is likely always to have residual Ca2+/CaM bound. This may be partially Ca2+-saturated CaM (19).Persistent autonomous activity conferred by Thr286 autophosphorylation is thought to enable αCaMKII to function as a memory molecule (20, 21). In contrast, however, following the development of chemical long term potentiation, rapid inactivation has also been reported (22). The extent of an autonomous activity is further obscured by the finding that Ca2+ sequestration induces a burst of autophosphorylation at residues Thr305/306, followed by a loss of activity (23). Moreover, when examined across a broad range of [Ca2+], the Ca2+/CaM dependence of phospho-Thr286-αCaMKII activity is apparent (19). It is thus vital to establish the mechanisms of activation and inactivation of αCaMKII at the molecular level in order to understand how it may function physiologically in learning and memory. To this end, it is necessary to dissect the mechanisms of Ca2+/CaM dissociation, Thr305/306 autophosphorylation, and inactivation of phospho-Thr286-αCaMKII and to establish the time window for autonomous Ca2+/CaM-independent activity.  相似文献   

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

8.
The Ca2+-binding protein calmodulin (CaM) has been shown to bind directly to cytoplasmic domains of some G protein-coupled receptors, including the dopamine D2 receptor. CaM binds to the N-terminal portion of the long third intracellular loop of the D2 receptor, within an Arg-rich epitope that is also involved in the binding to Gi/o proteins and to the adenosine A2A receptor, with the formation of A2A-D2 receptor heteromers. In the present work, by using proteomics and bioluminescence resonance energy transfer (BRET) techniques, we provide evidence for the binding of CaM to the A2A receptor. By using BRET and sequential resonance energy transfer techniques, evidence was obtained for CaM-A2A-D2 receptor oligomerization. BRET competition experiments indicated that, in the A2A-D2 receptor heteromer, CaM binds preferentially to a proximal C terminus epitope of the A2A receptor. Furthermore, Ca2+ was found to induce conformational changes in the CaM-A2A-D2 receptor oligomer and to selectively modulate A2A and D2 receptor-mediated MAPK signaling in the A2A-D2 receptor heteromer. These results may have implications for basal ganglia disorders, since A2A-D2 receptor heteromers are being considered as a target for anti-parkinsonian agents.G-protein-coupled receptors are able to form homo- and hetero-oligomers with unique biochemical and functional characteristics (17), and they are easily detected in vitro by using biophysical techniques (810). Heteromers of adenosine A2A and dopamine D2 receptors were one of the first G-protein-coupled receptor heteromers to be described (11). A close physical interaction between both receptors was shown using co-immunoprecipitation and co-localization assays (11) and fluorescence and bioluminescence resonance energy transfer (FRET2 or BRET) techniques (1214). At the biochemical level, two types of antagonistic A2A-D2 receptor interactions have been discovered that may explain the A2A-D2 receptor interactions described both at the neuronal and behavioral level (11, 1518). First, by means of an allosteric interaction in the receptor heteromer, stimulation of A2A receptor decreases the affinity of D2 receptor for their agonists (12). Second, the stimulation of the Gi/o-protein-coupled D2 receptor inhibits the cAMP accumulation induced by the stimulation of the Gs/olf-protein-coupled A2A receptor (11, 17, 18). In view of the well known role of dopamine in Parkinson disease, schizophrenia, and drug addiction, it has been suggested that the A2A-D2 receptor interactions in the central nervous system may provide new therapeutic approaches to combat these disorders (16, 19).An epitope-epitope electrostatic interaction between an Arg-rich epitope of the N terminus of the third intracellular loop (3IL) of the D2 receptor and an epitope containing a phosphorylated Ser localized in the distal part of the C terminus of the A2A receptor is involved in A2A-D2 receptor heteromer interface (14, 20, 21). The same Arg-rich epitope of the D2 receptor is able to interact with CaM (2225). In the absence of phosphorylated residues, adjacent aspartates or glutamates, which are abundant in CaM, may also form non-covalent complexes with Arg-rich epitopes (26). Therefore, CaM can potentially convey a Ca2+ signal to the D2 receptor through direct binding to the 3IL of the D2 receptor (22). Mass spectrometry data have shown that bovine CaM can form multiple non-covalent complexes with an Arg-rich peptide corresponding to the N-terminal region of the 3IL of the D2 receptor (VLRRRRKRVN) (24) as well as a peptide from the proximal C terminus of the A2A receptor (24). This epitope, whose sequence is 291RIREFRQTFR300 in the human A2A receptor, also contains several Arg residues. Since the suspected interaction between the A2A receptor and CaM was awaiting confirmation by assays using complete proteins, the present study was undertaken to demonstrate the existence of interactions between the A2A receptor and CaM both in a recombinant protein expression cell system and in the brain. A proteomics approach was used for the discovery of protein-protein interactions between the A2A receptor and CaM in rat brain, whereas BRET in transfected cells demonstrated a direct interaction between CaM and this receptor. Furthermore, by using BRET and sequential resonance energy transfer (SRET) techniques and analyzing MAPK signaling in transfected cells, evidence was obtained for CaM-A2A-D2 receptor oligomerization and a selective Ca2+-mediated modulation of A2A and D2 receptor function in the A2A-D2 receptor heteromer.  相似文献   

9.
We have investigated in detail the role of intra-organelle Ca2+ content during induction of apoptosis by the oxidant menadione while changing and monitoring the Ca2+ load of endoplasmic reticulum (ER), mitochondria, and acidic organelles. Menadione causes production of reactive oxygen species, induction of oxidative stress, and subsequently apoptosis. In both pancreatic acinar and pancreatic tumor AR42J cells, menadione was found to induce repetitive cytosolic Ca2+ responses because of the release of Ca2+ from both ER and acidic stores. Ca2+ responses to menadione were accompanied by elevation of Ca2+ in mitochondria, mitochondrial depolarization, and mitochondrial permeability transition pore (mPTP) opening. Emptying of both the ER and acidic Ca2+ stores did not necessarily prevent menadione-induced apoptosis. High mitochondrial Ca2+ at the time of menadione application was the major factor determining cell fate. However, if mitochondria were prevented from loading with Ca2+ with 10 μm RU360, then caspase-9 activation did not occur irrespective of the content of other Ca2+ stores. These results were confirmed by ratiometric measurements of intramitochondrial Ca2+ with pericam. We conclude that elevated Ca2+ in mitochondria is the crucial factor in determining whether cells undergo oxidative stress-induced apoptosis.Apoptosis, a mechanism of programmed cell death, usually occurs through intrinsic or extrinsic apoptotic pathways. The caspases involved in apoptosis can be split into two groups, the initiator caspases such as caspase-9 and effector caspases such as caspase-3. Effector caspases are activated by initiator caspases and mediate many of the morphological cellular changes associated with apoptosis (1).Calcium is an important signaling ion involved in the regulation of many physiological as well as pathological cellular responses (2). In the pancreas, we have shown that Ca2+ signals elicit enzyme secretion (3), apoptosis (46), and pathological intracellular activation of digestive enzymes (7). As such, there must be mechanisms in place by which the cell can differentiate between apoptotic and non-apoptotic Ca2+ signals.The spatiotemporal pattern of calcium signaling is crucial for the specificity of cellular responses. For example, repetitive cytosolic calcium spikes confined to the apical region of the pancreatic acinar cell are elicited by physiological stimulation with acetylcholine (ACh) or cholecystokinin (CCK) and result in physiological secretion of zymogen granules (8, 9). However, a sustained global increase in free cytosolic Ca2+ induced by supramaximal stimulation with CCK, which resembles prolonged hyperstimulation of pancreatic acinar cells in the pathophysiology of acute pancreatitis, can lead to premature activation of digestive enzymes and vacuole formation within the cell (1012). Alternatively, global repetitive calcium spikes induced in the pancreatic acinar cell in response to oxidant stress can lead to induction of the mitochondrial permeability transition pore (mPTP)4 and apoptosis (4, 5, 13).To understand the role of calcium in apoptosis, several investigators have examined the influence of intracellular stores on the molding of calcium signals that lead to cell death (1416). It has been well established in a range of cell types that the endoplasmic reticulum (ER) is the major intracellular calcium store required for induction of apoptosis. Pinton et al. (17) have shown that decreasing ER Ca2+ concentration with tBuBHQ increased HeLa cell survival in response to oxidant stress induced by ceramide. Scorrano and Korsmeyer (18) also observed that double knock-out Bax and Bak (pro-apoptotic proteins) mouse fibroblasts displayed a reduced resting concentration of ER Ca2+ compared with wild type and were resistant to induction of apoptosis by various stimulants, including ceramide. These important studies strongly suggest that the concentration of Ca2+ in the ER is a critical determinant of cellular susceptibility to apoptotic stimuli in the cell types studied.A key event in early apoptosis is permeabilization of the mitochondrial membrane. The mPTP is a pore whose molecular composition is still debated (19). Activation of an open pore state can result in swelling of the mitochondrial matrix and release of the apoptogenic proteins from the intermembrane space (20).One important activator of the mPTP is Ca2+ (2022), a function which implicates Ca2+ in the initiation of apoptosis (23, 24). Once Ca2+ is released from the ER into the cytoplasm, mitochondria take up part of the released Ca2+ to prevent propagation of large calcium waves (2527). This influx is followed by calcium efflux from the mitochondria back into the cytosol (28, 29). An increase in mitochondrial Ca2+ concentration in response to physiological stimuli induces increased activity of the mitochondrial respiratory chain and the synthesis of ATP to meet with increasing energy demands on the cell. When mitochondria are exposed to a pathological overload of calcium, opening of the mPTP is triggered, leading to mitochondrial dysfunction and eventually cell death. The mechanism through which calcium can trigger mPTP opening is still unclear and may involve cyclophilin D (30) and voltage-dependent anion channel (31). The mitochondria are endowed with selective and efficient calcium uptake (a calcium-selective uniporter) and release mechanisms (Ca2+/Na exchanger, Ca2+/H+ exchanger, and mPTP) (16, 29, 32, 33).Oxidant stress is a well known inducer of apoptosis in several cell types (34) and is thought to play an important role in the pathogenesis of acute pancreatitis (35). We have used the quinone compound menadione to induce oxidative stress in the pancreatic acinar cell. Menadione is metabolized by flavoprotein reductase to semiquinone and then is oxidized back to quinone, resulting in generation of superoxide anion radicals, hydrogen peroxide, and other reactive oxygen species (ROS) (36). In vivo, menadione causes depolarization and swelling of the mitochondria (37). In pancreatic acinar cells, treatment with menadione not only produces an increase in ROS, but has also been found to evoke cytosolic Ca2+ responses, mPTP opening, activation of caspases and apoptotic cell death (4, 5). When cells were pretreated with the calcium chelator BAPTA-AM, menadione was unable to induce apoptosis, indicating that oxidant stress-induced apoptosis in the pancreatic acinar cell is highly calcium-dependent. Here we show that in pancreatic acinar cells, oxidative stress-induced apoptosis is strongly dependent on the Ca2+ concentration within mitochondria at the time of ROS production.  相似文献   

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

11.
Intracellular Ca2+ mobilization plays an important role in a wide variety of cellular processes, and multiple second messengers are responsible for mediating intracellular Ca2+ changes. Here we explored the role of one endogenous Ca2+-mobilizing nucleotide, cyclic adenosine diphosphoribose (cADPR), in the proliferation and differentiation of neurosecretory PC12 cells. We found that cADPR induced Ca2+ release in PC12 cells and that CD38 is the main ADP-ribosyl cyclase responsible for the acetylcholine (ACh)-induced cADPR production in PC12 cells. In addition, the CD38/cADPR signaling pathway is shown to be required for the ACh-induced Ca2+ increase and cell proliferation. Inhibition of the pathway, on the other hand, accelerated nerve growth factor (NGF)-induced neuronal differentiation in PC12 cells. Conversely, overexpression of CD38 increased cell proliferation but delayed NGF-induced differentiation. Our data indicate that cADPR plays a dichotomic role in regulating proliferation and neuronal differentiation of PC12 cells.Mobilization of intracellular Ca2+ stores is involved in diverse cell functions, including fertilization, cell proliferation, and differentiation (14). At least three endogenous Ca2+-mobilizing messengers have been identified, including inositol trisphosphate (IP3),3 nicotinic adenine acid dinucleotide phosphate (NAADP), and cyclic adenosine diphosphoribose (cADPR). Similar to IP3, cADPR can mobilize calcium release in a wide variety of cell types and species, from protozoa to animals. The cADPR-mediated Ca2+ signaling has been indicated in a variety of cellular processes (57), from abscisic acid signaling and regulation of the circadian clock in plants, to mediating long-term synaptic depression in hippocampus.Ample evidence shows that the ryanodine receptors are the main intracellular targets for cADPR (1, 2, 8). Ryanodine receptors (RyRs) are intracellular Ca2+ channels widely expressed in various cells and tissues, including muscles and neurons. It is the major cellular mediator of Ca2+-induced Ca2+ release (CICR) in cells. There are three isoforms of ryanodine receptors: RyR1, RyR2, and RyR3, all of which have been implicated in the cADPR signaling (1, 2, 8). However, evidence regarding cADPR acting directly on the receptors is lacking (9). It has been suggested that accessory proteins, such as calmodulin and FK506-binding protein (FKBP), may be involved instead (1015).cADPR is formed from nicotinamide adenine dinucleotide (NAD) by ADP-ribosyl cyclases. Six ADP-ribosyl cyclases have been identified so far: Aplysia ADP-ribosyl cyclase, three sea urchin homologues (16, 17), and two mammalian homologues, CD38 and CD157 (18). CD38 is a membrane-bound protein and the main mammalian ADP-ribosyl cyclase. As a novel multifunctional enzyme, CD38 catalyzes the synthesis and hydrolysis of both cADPR and NAADP, two structurally and functionally distinct Ca2+ messengers. Virtually all mammalian tissues ever examined have been shown to express CD38. CD38 knock-out mice exhibit multiple physiological defects, ranging from impaired immune responses, metabolic disturbances, to behavioral modifications (1, 6, 18).CD38 was originally identified as a lymphocyte differentiation antigen (18). Indeed, CD38/cADPR has been linked to cell differentiation (5). For example, in human HL-60 cells, CD38 expression and the consequential accumulation of cADPR play a causal role in mediating granulocytic differentiation (19). In addition, expression of CD38 in HeLa and 3T3 cells not only increased intracellular Ca2+ concentration but also induced cell proliferation by significantly reducing the S phase duration, leading to shortened cell doubling time (20). The ability of cADPR to increase cell proliferation has also been observed in human T cells (21), human hemopoietic progenitors (22), human peripheral blood mononuclear cells (23), human mesenchymal stem cells (24), and murine mesangial cells (25).The PC12 cell line was derived from rat adrenal medulla and has been used extensively as a neuronal model, since it exhibits many of the functions observed in primary neuronal cultures (26). Most importantly, PC12 cells can be induced by nerve growth factor (NGF) to differentiate into cells with extensive neurite outgrowths, resembling neuronal dendritic trees (26, 27). In contrast to NGF, numerous growth factors and neurotransmitters can induce the proliferation of PC12 cells instead (26). Both IP3 receptor- and ryanodine receptor-mediated Ca2+ stores have been shown to be present in PC12 cells (2831). The type 2 ryanodine receptor is expressed in PC12 cells and activation of the NO/cGMP pathway in PC12 cells results in calcium mobilization, which is mediated by cADPR and similar to that seen in sea urchin eggs (32). It has been demonstrated that NAADP, another Ca2+-mobilizing messenger, is also a potent neuronal differentiation inducer in PC12 cells, while IP3 exhibits no such role (33, 34). Whether cADPR is involved in the proliferation and differentiation of PC12 cells is unknown.Here we show that activation of the CD38/cADPR/Ca2+ signaling is required for the ACh-induced proliferation in PC12 cells, while inhibition of the pathway accelerates NGF-induced neuronal differentiation. Our data indicate that cADPR is important in regulating cell proliferation and neuronal differentiation in PC12 cells.  相似文献   

12.
STIM1 and Orai1 have been reported to interact upon store depletion culminating in Ca2+ release-activated Ca2+ current activation. Recently, the essential region has been identified within the STIM1 C terminus that includes the second coiled-coil domain C-terminally extended by ∼50 amino acids and exhibits a strong binding to the Orai1 C terminus. Based on the homology within the Orai family, an analogous scenario might be assumed for Orai2 as well as Orai3 channels as both are activated in a similar STIM1-dependent manner. A combined approach of electrophysiology and Foerster resonance energy transfer microscopy uncovered a general mechanism in the communication of STIM1 with Orai proteins that involved the conserved putative coiled-coil domains in the respective Orai C terminus and the second coiled-coil motif in the STIM1 C terminus. A coiled-coil single mutation in the Orai1 C terminus abrogated communication with the STIM1 C terminus, whereas an analogous mutation in Orai2 and Orai3 still allowed for their moderate activation. However, increasing coiled-coil probability by a gain of function deletion in Orai1 or by generating an Orai1-Orai3 chimera containing the Orai3 C terminus recovered stimulation to a similar extent as with Orai2/3. At the level of STIM1, decreasing probability of the second coiled-coil domain by a single mutation within the STIM1 C terminus abolished activation of Orai1 but still enabled partial stimulation of Orai2/3 channels. A double mutation within the second coiled-coil motif of the STIM1 C terminus fully disrupted communication with all three Orai channels. In aggregate, the impairment in the overall communication between STIM1 and Orai channels upon decreasing probabilities of either one of the putative coiled-coil domains in the C termini might be compatible with the concept of their functional, heteromeric interaction.Store-operated Ca2+ entry is a key to cellular regulation of short term responses such as contraction and secretion as well as long term processes like proliferation and cell growth (1). The prototypic and best characterized store-operated channel is the Ca2+ release-activated Ca2+ (CRAC)5 channel (26). However, its molecular components have remained elusive until 4 years ago; the STIM1 (stromal interacting molecule 1) (7, 8) and later on Orai1 (911) have been identified as the two limiting components for CRAC activation. STIM1 is an ER-located Ca2+ sensor, and store depletion triggers its aggregation into punctae close to the plasma membrane, resulting in stimulation of CRAC currents (12, 13). Its N terminus is located in the ER lumen and contains an EF-hand Ca2+-binding motif, which senses the ER Ca2+ level, and a sterile α-motif, which is suggested to mediate homomeric STIM1 aggregation (1416). In the cytosolic STIM1 C terminus, two coiled-coil regions overlapping with the ezrin-radixin-moesin-like domain and a lysine-rich region are essential for CRAC activation (14, 17, 18). Three recent studies have independently identified the ezrin-radixin-moesin domain as the essential Orai activating domain, named SOAR (STIM1 Orai-activating region) (20) which represents so far the shortest active fragment, OASF (Orai-activating small fragment) (21) or CAD (CRAC-activating domain) (22), which includes the second, more C terminally located coiled-coil domain and the following ∼55 amino acids. The latter amino acids are suggested to contain an additional cytosolic homomerization domain indispensable for OASF homomerization and Orai activation (21).The Orai family includes three highly Ca2+-selective ion channels (Orai1–3) that locate to the plasma membrane, and each protein contains four predicted transmembrane segments with cytosolic N and C termini (10). All three Orai proteins possess a conserved putative coiled-coil domain in the C terminus (23, 24), whereas only the N terminus of Orai1 consists of a proline/arginine-rich region (25). Orai1 has been assumed to act in concert with STIM1 (10, 27)-activating inward Ca2+ currents after store depletion. The two other members of the Orai family, Orai2 and Orai3, display similar but smaller store-operated inward Ca2+ currents when co-expressed with STIM1 with distinct inactivation profiles, permeability properties, and 2-aminoethoxydiphenyl borate sensitivity (2832). Recently, we have provided evidence for a store depletion-induced, dynamic coupling of STIM1 to Orai1 that involves the putative coiled-coil domain in the C terminus of Orai1 (33). Furthermore, the C terminus of STIM1, in particular the essential cytosolic region 344–442 as narrowed down by SOAR, OASF, and CAD (2022), has been established as the key fragment for CRAC as well as Orai1 activation, because its expression alone, without the necessity to deplete ER store, is sufficient for constitutive current activation (18, 32, 33). These fragments SOAR, OASF, and CAD when co-expressed with Orai1 (2022) exhibit enhanced plasma membrane localization in comparison with the complete STIM1 C terminus in the presence of Orai1. Specificity of interaction of SOAR to the Orai1 C terminus has been shown by its disruption (20) employing the Orai1 L273S mutant (33). Park et al. (22) have provided additional, conclusive evidence for a direct binding by combining multiple biochemical approaches demonstrating CAD interaction with Orai1.This study focused specifically on the role of the putative coiled-coil domains of STIM1 as well as Orai proteins in their coupling. Coiled-coils generally function as protein-protein interaction sites with the ability of dynamic protein assembly and disassembly (3537). We suggest the C-terminal, putative coiled-coil domains in all three Orai proteins and the second coiled-coil motif of STIM1 as essential for STIM1/Orai communication. Moreover, the single point coiled-coil STIM1 L373S mutant allowed for differential activation of Orai channels partially stimulating Orai2 as well as Orai3 but not Orai1.  相似文献   

13.
14.
15.
16.
The binding of the adaptor protein APPL1 to adiponectin receptors is necessary for adiponectin-induced AMP-activated protein kinase (AMPK) activation in muscle, yet the underlying molecular mechanism remains unknown. Here we show that in muscle cells adiponectin and metformin induce AMPK activation by promoting APPL1-dependent LKB1 cytosolic translocation. APPL1 mediates adiponectin signaling by directly interacting with adiponectin receptors and enhances LKB1 cytosolic localization by anchoring this kinase in the cytosol. Adiponectin also activates another AMPK upstream kinase Ca2+/calmodulin-dependent protein kinase kinase by activating phospholipase C and subsequently inducing Ca2+ release from the endoplasmic reticulum, which plays a minor role in AMPK activation. Our results show that in muscle cells adiponectin is able to activate AMPK via two distinct mechanisms as follows: a major pathway (the APPL1/LKB1-dependent pathway) that promotes the cytosolic localization of LKB1 and a minor pathway (the phospholipase C/Ca2+/Ca2+/calmodulin-dependent protein kinase kinase-dependent pathway) that stimulates Ca2+ release from intracellular stores.Adiponectin, an adipokine abundantly expressed in adipose tissue, exhibits anti-diabetic, anti-inflammatory, and anti-atherogenic properties and hence is a potential therapeutic target for various metabolic diseases (13). The beneficial effects of adiponectin are mediated through the direct interaction of adiponectin with its cell surface receptors, AdipoR1 and AdipoR2 (4, 5). Adiponectin increases fatty acid oxidation and glucose uptake in muscle cells by activating AMP-activated protein kinase (AMPK)3 (4, 6), which depends on the interaction of AdipoR1 with the adaptor protein APPL1 (Adaptor protein containing Pleckstrin homology domain, Phosphotyrosine binding domain, and Leucine zipper motif) (5). However, the underlying mechanisms by which APPL1 mediates adiponectin signaling to AMPK activation and other downstream targets remain unclear.AMPK is a serine/threonine protein kinase that acts as a master sensor of cellular energy balance in mammalian cells by regulating glucose and lipid metabolism (7, 8). AMPK is composed of a catalytic α subunit and two noncatalytic regulatory subunits, β and γ. The NH2-terminal catalytic domain of the AMPKα subunit is highly conserved and contains the activating phosphorylation site (Thr172) (9). Two AMPK variants, α1 and α2, exist in mammalian cells that show different localization patterns. AMPKα1 subunit is localized in non-nuclear fractions, whereas the AMPKα2 subunit is found in both nucleus and non-nuclear fractions (10). Biochemical regulation of AMPK activation occurs through various mechanisms. An increase in AMP level stimulates the binding of AMP to the γ subunit, which induces a conformational change in the AMPK heterotrimer and results in AMPK activation (11). Studies have shown that the increase in AMPK activity is not solely via AMP-dependent conformational change, rather via phosphorylation by upstream kinases, LKB1 and CaMKK. Dephosphorylation by protein phosphatases is also important in regulating the activity of AMPK (12).LKB1 has been considered as a constitutively active serine/threonine protein kinase that is ubiquitously expressed in all tissues (13, 14). Under conditions of high cellular energy stress, LKB1 acts as the primary AMPK kinase through an AMP-dependent mechanism (1517). Under normal physiological conditions, LKB1 is predominantly localized in the nucleus. LKB1 is translocated to the cytosol, either by forming a heterotrimeric complex with Ste20-related adaptor protein (STRADα/β) and mouse protein 25 (MO25α/β) or by associating with an LKB1-interacting protein (LIP1), to exert its biological function (1822). Although LKB1 has been shown to mediate contraction- and adiponectin-induced activation of AMPK in muscle cells, the underlying molecular mechanisms remain elusive (15, 23).CaMKK is another upstream kinase of AMPK, which shows considerable sequence and structural homology with LKB1 (2426). The two isoforms of CaMKK, CaMKKα and CaMKKβ, encoded by two distinct genes, share ∼70% homology at the amino acid sequence level and exhibit a wide expression in rodent tissues, including skeletal muscle (2734). Unlike LKB1, AMPK phosphorylation mediated by CaMKKs is independent of AMP and is dependent only on Ca2+/calmodulin (35). Hence, it is possible that an LKB1-independent activation of AMPK by CaMKK exists in muscle cells. However, whether and how adiponectin stimulates this pathway in muscle cells are not known.In this study, we demonstrate that in muscle cells adiponectin induces an APPL1-dependent LKB1 translocation from the nucleus to the cytosol, leading to increased AMPK activation. Adiponectin also activates CaMKK by stimulating intracellular Ca2+ release via the PLC-dependent mechanism, which plays a minor role in activation of AMPK. Taken together, our results demonstrate that enhanced cytosolic localization of LKB1 and Ca2+-induced activation of CaMKK are the mechanisms underlying adiponectin-stimulated AMPK activation in muscle cells.  相似文献   

17.
18.
19.
Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.  相似文献   

20.
Data, both for and against the presence of a mitochondrial nitric-oxide synthase (NOS) isoform, is in the refereed literature. However, irrefutable evidence has not been forthcoming. In light of this controversy, we designed studies to investigate the existence of the putative mitochondrial NOS. Using repeated differential centrifugation followed by Percoll gradient fractionation, ultrapure, never frozen rat liver mitochondria and submitochondrial particles were obtained. Following trypsin digestion and desalting, the mitochondrial samples were analyzed by nano-HPLC-coupled linear ion trap-mass spectrometry. Linear ion trap-mass spectrometry analyses of rat liver mitochondria as well as submitochondrial particles were negative for any peptide from any NOS isoform. However, recombinant neuronal NOS-derived peptides from spiked mitochondrial samples were easily detected, down to 50 fmol on column. The protein calmodulin (CaM), absolutely required for NOS activity, was absent, whereas peptides from CaM-spiked samples were detected. Also, l-[14C]arginine to l-[14C]citrulline conversion assays were negative for NOS activity. Finally, Western blot analyses of rat liver mitochondria, using NOS (neuronal or endothelial) and CaM antibodies, were negative for any NOS isoform or CaM. In conclusion, and in light of our present limits of detection, data from carefully conducted, properly controlled experiments for NOS detection, utilizing three independent yet complementary methodologies, independently as well as collectively, refute the claim that a NOS isoform exists within rat liver mitochondria.Nitric oxide (NO·)2 is a highly diffusible, hydrophobic, and gaseous free radical (1) that is responsible for autocrine and paracrine signaling activities (2). NO· can readily partition into and through membranes (35) to influence biological functions such as blood pressure regulation, platelet aggregation and adhesion, neurotransmission, and cellular defense (4, 611). The mechanism by which NO· influences biological functions is by binding to target proteins that contain heme and/or thiol(s). Alternatively, NO· can combine with to produce the highly reactive species peroxynitrite.Mitochondria are highly compartmentalized, membranous organelles that contain abundant amounts of reactive hemoproteins and thiols (12, 13), to which NO· may bind reversibly (14, 15) or irreversibly (1618). Mitochondria also generate various amounts of during the process of cellular respiration (19, 20). Studies conducted during the past decade have suggested that NO· can diffuse into mitochondria and cause mitochondrial dysfunction by reversibly inhibiting cytochrome c oxidase (14, 21, 22) and NADH dehydrogenase (23).In the mid-90s, a putative variant of NOS was proposed to reside within mitochondria. Initially, Kobzik et al. (24) and Hellsten and co-workers (25) observed an apparent endothelial NOS (eNOS) immunoreactivity in skeletal muscle mitochondria. Simultaneously, Bates et al. (26, 27) observed an apparent eNOS histochemical reactivity in inner mitochondrial membrane preparations, isolated from rat liver, brain, heart, skeletal muscle, and kidney. Tatoyan and Giulivi (28), acting on these initial observations, performed experiments in an attempt to confirm the identity of this putative mtNOS. Relying on immunochemical analysis, Tatoyan and Giulivi (28) claimed that inducible NOS (iNOS) was the NOS isoform present in rat liver mitochondria. This same group using mass spectrometry later presented data in support of the putative mtNOS being a variant of nNOS (29). Ghafourifar and Richter (30) had reported previously that the putative mtNOS was calcium-sensitive and constitutive in nature. Since these reports, different groups have reported the presence of each of the three main isoforms of NOS within mitochondria (29, 31, 32). Also, biochemical characterization of the putative mtNOS performed by Giulivi and co-workers (29) revealed certain post-translational modifications (myristoylation and phosphorylation of the protein) that are thought to be unique to eNOS. During the last decade, various reports have supported the presence of at least one of the three main isoforms of NOS residing in mitochondria. However, the more recent reports tend to question this claim (3336). Because of the contradictory reports regarding the existence of a putative mtNOS, Brookes (33) compiled a critical and thorough review of the literature published up to 2003 dealing with the putative mtNOS. This review brought to light the diverse technical issues involved in the aforementioned studies. Major issues were the degree of purity of mitochondrial preparations (37, 38), shortcomings of measurement methodology (29, 3941), use of inappropriate, or total lack of, experimental controls and confusing technical practices. Lacza et al. (42) has reviewed the more recent developments in the area of mitochondrial NO· production and discussed some of the shortcomings of certain techniques still being used.In light of this ongoing controversy regarding the presence or absence of a mtNOS, we designed and carefully conducted properly controlled studies to either confirm or refute the existence of any NOS isoform within mitochondria. Ultrapure rat liver mitochondria were isolated using repeated differential centrifugation followed by Percoll gradient purification. Proteomic analyses were then performed using a nano-HPLC-coupled nanospray LTQ-MS. To avoid the interfering factors that are rampant in NO· trapping assays (43), the NOS-catalyzed conversion of l-[14C]arginine to l-[14C]citrulline was used to probe for NOS activity in mitochondria. Appropriate controls were employed and, for inhibition studies, high concentrations of l-thiocitrulline (TC) (44) were used. Additionally, immunochemical analyses were performed with ultrapure mitochondria using nNOS, eNOS, and CaM antibodies. The problems faced with the commonly used techniques in mtNOS studies are discussed.  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