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 582 毫秒
1.
2.
3.
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   

4.
Nine membrane-bound mammalian adenylyl cyclases (ACs) have been identified. Type 1 and 8 ACs (AC1 and AC8), which are both expressed in the brain and are stimulated by Ca2+/calmodulin (CaM), have discrete neuronal functions. Although the Ca2+ sensitivity of AC1 is higher than that of AC8, precisely how these two ACs are regulated by Ca2+/CaM remains elusive, and the basis for their diverse physiological roles is quite unknown. Distinct localization of the CaM binding domains within the two enzymes may be essential to differential regulation of the ACs by Ca2+/CaM. In this study we compare in detail the regulation of AC1 and AC8 by Ca2+/CaM both in vivo and in vitro and explore the different role of each Ca2+-binding lobe of CaM in regulating the two enzymes. We also assess the relative dependence of AC1 and AC8 on capacitative Ca2+ entry. Finally, in real-time fluorescence resonance energy transfer-based imaging experiments, we examine the effects of dynamic Ca2+ events on the production of cAMP in cells expressing AC1 and AC8. Our data demonstrate distinct patterns of regulation and Ca2+ dependence of AC1 and AC8, which seems to emanate from their mode of regulation by CaM. Such distinctive properties may contribute significantly to the divergent physiological roles in which these ACs have been implicated.Nine membrane-bound mammalian adenylyl cyclases (ACs),2 AC1–AC9, have been identified (1). They possess a common predicted structure (2)3 and are stimulated by forskolin (FSK; except AC9) and G, although they are distributed and regulated differently (1, 3, 4). Four ACs are regulated by physiological concentrations of Ca2+ and thereby provide a critical link between the Ca2+- and cAMP-signaling pathways (3, 5); AC5 and AC6 are directly inhibited by Ca2+, whereas AC1 and AC8 are stimulated by Ca2+ in a calmodulin (CaM)-dependent manner (5). AC3 is also regulated by CaM in vitro, although this requires supramicromolar concentration of Ca2+ (6), and in vivo AC3 is inhibited by Ca2+ via CaM kinase II (7), unlike AC1 and AC8.AC1 is closely related in sequence to the Ca2+/CaM-stimulable rutabaga AC from Drosophila, which is important in Drosophila learning tasks (810). AC1 and the other Ca2+/CaM-stimulable mammalian AC, AC8, have also been implicated in learning and memory (11). As a means of establishing their proposed roles, single and/or double AC1 and AC8 knockout mice have been generated. Mouse models have demonstrated that Ca2+/CaM-stimulable ACs are involved in long-term potentiation and long-term memory (12). However, despite the general view that AC1 and AC8 can behave similarly, discrete physiological actions of each isoform are becoming apparent. Recent studies by Zhuo''s group demonstrated that AC1 specifically participates in N-methyl-d-aspartic acid receptor-induced neuronal excitotoxicity (13) and an increase in GluR1 synthesis induced by blocking AMPA receptors (14). Furthermore, Nicol and colleagues (15, 16) showed a contribution of AC1, but not AC8, in axon terminal refinement in the retina. On the other hand, AC8 specifically was seen to be responsible for retrieval from adaptive presynaptic silencing (17) and the acquiring of new spatial information (18). These differences in physiological roles must reflect not only differences in their distributions but also presumably in their regulatory properties. Both enzymes are expressed in brain; AC1 is neuro-specific, whereas the expression of AC8 is more widespread (1, 12). Within the central nervous system, AC1 is abundant in the hippocampus, the cerebral cortex, and the granule cells of the cerebellum, whereas AC8 has a high expression level in the thalamus and the cerebral cortex (19). Studies of mouse brain revealed that AC1 is distributed pre-synaptically and AC8 post-synaptically (18, 20).Although physiological differences in the roles of these two enzymes are suggested from the studies outlined above, the regulatory mechanisms that might underlie these differences are not. AC1 is more sensitive to Ca2+ than is AC8 in vitro (21, 22), yet details on how these two enzymes are regulated by Ca2+/CaM are sparse. In non-excitable cells, a Ca2+ elevation caused by capacitative Ca2+ entry (CCE), the mode of Ca2+ entry triggered by emptying Ca2+ from internal stores (23), preferentially stimulates AC1 and AC8 (21). Although stimulation of AC8 by CCE has been shown to be at least partially dependent on its localization at lipid rafts (24), whether AC1 is also targeted to this region of plasma membranes has never been addressed. In addition, CaM regulation of AC1 and AC8 has not been compared in detail, although CaM appears to bind to different domains of the two enzymes. AC8 utilizes two CaM binding domains: a classic amphipathic “1-5-8-14” motif at the N terminus and an IQ-like motif in the C2b domain (25). A recent study indicates that CaM pre-associates with the N terminus of AC8, where it becomes fully saturated upon a Ca2+ rise, and activates the enzyme via a C-terminally mediated relief of auto-inhibitory mechanisms (26). By contrast, only residues 495–522 of the C1b region of AC1 have been shown to bind CaM in a Ca2+-dependent manner (27, 28). With the presence of only one CaM binding domain in AC1, a simpler mechanism of CaM regulation might be expected.CaM mediates the regulation of numerous Ca2+-dependent processes in eukaryotic cells (29). The protein possesses N- and C-terminal lobes, both of which contain two Ca2+ binding EF hands (EF1 and EF2 in the N lobe, and EF3 and EF4 in the C lobe (30)). Mutations in the EF hands have been valuable for investigating the interaction of CaM with its targets. Alanine substitutions in the EF12 pair or EF34 pair have generated CaM12 and CaM34 to investigate the independent function of the C and N lobes of CaM, respectively (31, 32).Against the background of the distinct physiological roles carried out by AC1 and AC8, we performed a detailed comparison of the two enzymes expressed in HEK 293 cells. Their sensitivity to Ca2+/CaM was compared both in vitro and in vivo; the possibility that they might be expressed in different domains of the plasma membrane was addressed; and putative lobe-specific CaM regulation was assessed using Ca2+-binding mutants of CaM. Single cell measurements using a FRET-based cAMP sensor were performed to compare the kinetic responses of the enzymes to physiological elevations of [Ca2+]i. The results demonstrate superficial similarities in the regulation of AC1 and AC8 but critical disparities in their mechanism of activation by the lobes of CaM and in the speed and pattern of their responsiveness to [Ca2+]i. These discrete behaviors provide a physiological opportunity for different outcomes to elevation of [Ca2+]i, depending on the AC that is expressed in particular contexts.  相似文献   

5.
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   

6.
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   

7.
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   

8.
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   

9.
10.
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   

11.
We have determined the structure of the sarco(endo)plasmic reticulum Ca2+-ATPase (SERCA) in an E2·Pi-like form stabilized as a complex with , an ATP analog, adenosine 5′-(β,γ-methylene)triphosphate (AMPPCP), and cyclopiazonic acid (CPA). The structure determined at 2.5Å resolution leads to a significantly revised model of CPA binding when compared with earlier reports. It shows that a divalent metal ion is required for CPA binding through coordination of the tetramic acid moiety at a characteristic kink of the M1 helix found in all P-type ATPase structures, which is expected to be part of the cytoplasmic cation access pathway. Our model is consistent with the biochemical data on CPA function and provides new measures in structure-based drug design targeting Ca2+-ATPases, e.g. from pathogens. We also present an extended structural basis of ATP modulation pinpointing key residues at or near the ATP binding site. A structural comparison to the Na+,K+-ATPase reveals that the Phe93 side chain occupies the equivalent binding pocket of the CPA site in SERCA, suggesting an important role of this residue in stabilization of the potassium-occluded E2 state of Na+,K+-ATPase.The Ca2+-ATPase from sarco(endo)plasmic reticulum of rabbit skeletal muscle (SERCA,5 isoform 1a) is a thoroughly studied member of the P-type ATPase family (1). SERCA possesses 10 transmembrane helices (M1 through M10) with both the N terminus and the C terminus facing the cytoplasmic side and three cytoplasmic domains, inserted in loops between M2 and M3 (A-domain) and between M4 and M5 (P- and N-domain) (2). The enzyme mediates the uptake of Ca2+ ions into the lumen of the sarcoplasmic reticulum (SR) after their release into the cytoplasm through calcium release channels during muscle contraction (3). SERCA, plasma membrane Ca2+-ATPase, and a third, Golgi-located secretory pathway Ca2+-ATPase are important factors in calcium and manganese homeostasis, transport, signaling, and regulation (4, 5).Crystal structures of all major states in the reaction cycle of SERCA have been determined. These include the Ca2E1·ATP state (6, 7) with high affinity Ca2+ binding sites accessible from the cytoplasmic side of the SR membrane, the calcium-occluded transition state (6), the open E2P state with luminal facing ion binding sites that have low affinity for Ca2+ and high affinity for protons (8) and the proton-occluded H2–3E2[ATP] state with a bound modulatory ATP (9). This considerable amount of structural information has turned the Ca2+-ATPase into a valuable model system for studies on structural rearrangements that take place during the catalytic cycle of P-type ATPases. SERCA is considered a promising drug target in medical research, with a particular focus on prostate cancer and infectious diseases. Several compounds have already been shown to bind and inhibit SERCA by stabilizing the enzyme in a particular conformational state. Thapsigargin (TG), cyclopiazonic acid (CPA), and 2,5-di-(tert-butyl) hydroquinone (BHQ) stabilize an E2-like state, and 1,3-dibromo-2,4,6-tri (methylisothiouronium)benzene stabilizes an E1-P-like conformation (1013). CPA is a toxic indole tetramic acid first isolated from Penicillium cyclopium (14) and later found to be produced by Aspergillus versicolor and Aspergillus flavus. Like TG, CPA specifically binds to and inhibits SERCA with nanomolar affinity (15). Indeed, CPA is widely used in biochemical and physiological studies on Ca2+ signaling and muscle function, where it causes Ca2+ store depletion due to specific inhibition of Ca2+ reuptake by SERCA. CPA and TG were originally proposed to bind to similar sites on SERCA (16), but recent crystal structures have shown a distinct site of interaction (17, 18). Despite these structural insights, a previously demonstrated magnesium dependence of CPA binding (19) remained unexplained, and opposing CPA binding modes were observed (see below).Tetramic acids are synthesized naturally, and more than 150 natural derivatives have been isolated from bacterial and fungal species (reviewed in Ref. 20). Tetramic acids possessing a 3-acyl group have the ability to chelate divalent metal ions. For instance, tenuazonic acid from the fungus Phoma sorghina has been shown to form complexes with Ca2+ and Mg2+ (21), as well as heavier metals such as Cu(II), Ni(II), and Fe(III) (22).Previously published crystallographic structures of the SERCA·CPA complex (PDB ID 2O9J and 2EAS) demonstrated that CPA binds within the proposed calcium access channel of SERCA. However, the structures did not reveal a role for magnesium, and the orientation of CPA within this binding site differed in the two studies (17, 18). To address these ambiguities, we have determined the crystal structure of SERCA in complex with , AMPPCP (an ATP analog), and Mn2+·CPA. The structure reveals novel insight into CPA binding, which we find to be mediated by a divalent cation, as demonstrated by means of the anomalous scattering properties of Mn2+. Further and improved refinement using previously deposited data (PDB ID 2O9J and 2OA0), in light of our new findings, also revealed a strong plausibility for a magnesium ion bound at this site. Furthermore, we find a new configuration of the bound AMPPCP nucleotide, addressing the modulatory role of ATP binding to the E2·Pi occluded conformation of SERCA.  相似文献   

12.
The psychostimulants d-amphetamine (AMPH) and methamphetamine (METH) release excess dopamine (DA) into the synaptic clefts of dopaminergic neurons. Abnormal DA release is thought to occur by reverse transport through the DA transporter (DAT), and it is believed to underlie the severe behavioral effects of these drugs. Here we compare structurally similar AMPH and METH on DAT function in a heterologous expression system and in an animal model. In the in vitro expression system, DAT-mediated whole-cell currents were greater for METH stimulation than for AMPH. At the same voltage and concentration, METH released five times more DA than AMPH and did so at physiological membrane potentials. At maximally effective concentrations, METH released twice as much [Ca2+]i from internal stores compared with AMPH. [Ca2+]i responses to both drugs were independent of membrane voltage but inhibited by DAT antagonists. Intact phosphorylation sites in the N-terminal domain of DAT were required for the AMPH- and METH-induced increase in [Ca2+]i and for the enhanced effects of METH on [Ca2+]i elevation. Calmodulin-dependent protein kinase II and protein kinase C inhibitors alone or in combination also blocked AMPH- or METH-induced Ca2+ responses. Finally, in the rat nucleus accumbens, in vivo voltammetry showed that systemic application of METH inhibited DAT-mediated DA clearance more efficiently than AMPH, resulting in excess external DA. Together these data demonstrate that METH has a stronger effect on DAT-mediated cell physiology than AMPH, which may contribute to the euphoric and addictive properties of METH compared with AMPH.The dopamine transporter (DAT)3 is a main target for psychostimulants, such as d-amphetamine (AMPH), methamphetamine (METH), cocaine (COC), and methylphenidate (Ritalin®). DAT is the major clearance mechanism for synaptic dopamine (DA) (1) and thereby regulates the strength and duration of dopaminergic signaling. AMPH and METH are substrates for DAT and competitively inhibit DA uptake (2, 3) and release DA through reverse transport (49). AMPH- and METH-induced elevations in extracellular DA result in complex neurochemical changes and profound psychiatric effects (2, 1016). Despite their structural and pharmacokinetic similarities, a recent National Institute on Drug Abuse report describes METH as a more potent stimulant than AMPH with longer lasting effects at comparable doses (17). Although the route of METH administration and its availability must contribute to the almost four times higher lifetime nonmedical use of METH compared with AMPH (18), there may also be differences in the mechanisms that underlie the actions of these two drugs on the dopamine transporter.Recent studies by Joyce et al. (19) have shown that compared with d-AMPH alone, the combination of d- and l-AMPH in Adderall® significantly prolonged the time course of extracellular DA in vivo. These experiments demonstrate that subtle structural features of AMPH, such as chirality, can affect its action on dopamine transporters. Here we investigate whether METH, a more lipophilic analog of AMPH, affects DAT differently than AMPH, particularly in regard to stimulated DA efflux.METH and AMPH have been reported as equally effective in increasing extracellular DA levels in rodent dorsal striatum (dSTR), nucleus accumbens (NAc) (10, 14, 20), striatal synaptosomes, and DAT-expressing cells in vitro (3, 6). John and Jones (21), however, have recently shown in mouse striatal and substantia nigra slices, that AMPH is a more potent inhibitor of DA uptake than METH. On the other hand, in synaptosomes METH inhibits DA uptake three times more effectively than AMPH (14), and in DAT-expressing COS-7 cells, METH releases DA more potently than AMPH (EC50 = 0.2 μm for METH versus EC50 = 1.7 μm for AMPH) (5). However, these differences do not hold up under all conditions. For example, in a study utilizing C6 cells, the disparity between AMPH and METH was not found (12).The variations in AMPH and METH data extend to animal models. AMPH- and METH-mediated behavior has been reported as similar (22), lower (20), or higher (23) for AMPH compared with METH. Furthermore, although the maximal locomotor activation response was less for METH than for AMPH at a lower dose (2 mg/kg, intraperitoneal), both drugs decreased locomotor activity at a higher dose (4 mg/kg) (20). In contrast, in the presence of a salient stimuli, METH is more potent in increasing the overall magnitude of locomotor activity in rats yet is equipotent with AMPH in the absence of these stimuli (23).The simultaneous regulation of DA uptake and efflux by DAT substrates such as AMPH and METH, as well as the voltage dependence of DAT (24), may confound the interpretation of existing data describing the action of these drugs. Our biophysical approaches allowed us to significantly decrease the contribution of DA uptake and more accurately determine DAT-mediated DA efflux with millisecond time resolution. We have thus exploited time-resolved, whole-cell voltage clamp in combination with in vitro and in vivo microamperometry and Ca2+ imaging to compare the impact of METH and AMPH on DAT function and determine the consequence of these interactions on cell physiology.We find that near the resting potential, METH is more effective than AMPH in stimulating DAT to release DA. In addition, at efficacious concentrations METH generates more current, greater DA efflux, and higher Ca2+ release from internal stores than AMPH. Both METH-induced or the lesser AMPH-induced increase in intracellular Ca2+ are independent of membrane potential. The additional Ca2+ response induced by METH requires intact phosphorylation sites in the N-terminal domain of DAT. Finally, our in vivo voltammetry data indicate that METH inhibits clearance of locally applied DA more effectively than AMPH in the rat nucleus accumbens, which plays an important role in reward and addiction, but not in the dorsal striatum, which is involved in a variety of cognitive functions. Taken together these data imply that AMPH and METH have distinguishable effects on DAT that can be shown both at the molecular level and in vivo, and are likely to be implicated in the relative euphoric and addictive properties of these two psychostimulants.  相似文献   

13.
Chlamydiales and Rickettsiales as metabolically impaired, intracellular pathogenic bacteria essentially rely on “energy parasitism” by the help of nucleotide transporters (NTTs). Also in plant plastids NTT-type carriers catalyze ATP/ADP exchange to fuel metabolic processes. The uptake of ATP4-, followed by energy consumption and the release of ADP3-, would lead to a metabolically disadvantageous accumulation of negative charges in form of inorganic phosphate (Pi) in the bacterium or organelle if no interacting Pi export system exists. We identified that Pi is a third substrate of several NTT-type ATP/ADP transporters. During adenine nucleotide hetero-exchange, Pi is cotransported with ADP in a one-to-one stoichiometry. Additionally, Pi can be transported in exchange with solely Pi. This Pi homo-exchange depends on the presence of ADP and provides a first indication for only one binding center involved in import and export. Furthermore, analyses of mutant proteins revealed that Pi interacts with the same amino acid residue as the γ-phosphate of ATP. Import of ATP in exchange with ADP plus Pi is obviously an efficient way to couple energy provision with the export of the two metabolic products (ADP plus Pi) and to maintain cellular phosphate homeostasis in intracellular living “energy parasites” and plant plastids. The additional Pi transport capacity of NTT-type ATP/ADP transporters makes the existence of an interacting Pi exporter dispensable and might explain why a corresponding protein so far has not been identified.Most organisms possess the capacity to resynthesize the fundamental energy currency ATP by fusion of ADP and Pi. Generally, in eukaryotes the major part of energy is produced in specialized organelles, the mitochondria. Mitochondrial ADP/ATP carriers (AACs)2 mediate the export of newly synthesized ATP in strict counter-exchange with cytosolic ADP and therefore provide energy to the cellular metabolism (1). Plants additionally generate high amounts of ATP during photosynthesis in chloroplasts. However, under conditions of limiting or missing photosynthetic activity, plant plastids depend on external energy supply (24). Specific nucleotide transporters (NTTs) located in the inner plastid envelope membrane mediate the required energy import (5). These transporters structurally, functionally, and phylogenetically differ from mitochondrial AACs. They catalyze the import of cytosolic ATP in exchange with stromal ADP, are monomers consisting of 12 predicted transmembrane helices, and are related to the functionally heterogeneous group of bacterial NTTs (5).Although most prokaryotic organisms are able to regenerate ATP and therefore are considered as energetically self-sustaining, the obligate intracellular living bacterial orders Chlamydiales and Rickettsiales are impaired in energy and nucleotide synthesis or even completely lost the corresponding pathways (68). Therefore, these bacteria, which comprise important human pathogens (9, 10), essentially rely on nucleotide and energy import. Bacterial NTTs catalyze the required import of a broad range of nucleotides and NAD or facilitate the counter-exchange of ATP and ADP (5, 1115). The latter process has been termed “energy parasitism” and obviously is of high importance for the survival of rickettsial and chlamydial cells (5, 1618).Although import measurements on intact Escherichia coli cells expressing the corresponding proteins allowed characterization of many bacterial and plastidial NTTs (1215, 1924), a very important physiological question is still not clarified. The uptake of ATP4- in exchange with ADP3- in absence of a concerted Pi export would result in a charge difference and a phosphate imbalance in the bacterial cell. In mitochondria, phosphate carriers metabolically cooperate with AACs because they provide Pi for ATP synthesis (25). Similarly, it was assumed that NTT-type ATP/ADP transporters cooperate with phosphate exporters to guarantee phosphate homeostasis in the bacterium or plastid. However, a Pi exporter interacting with ATP/ADP transporters is not known in “energy parasites” or plant plastids. Bacterial and plant phosphate transport systems rather facilitate Pi import or the counter-exchange of Pi and phosphorylated compounds and therefore do not allow net Pi export (2629). Furthermore, the newly identified plastidial (proton-driven) phosphate transporters are not preferentially expressed under conditions or in tissues that require ATP provision to the plastid (30, 31).Recently, we succeeded in the purification of the first recombinant NTT from Protochlamydia amoebophila (PamNTT1), a parachlamydial endosymbiont of the protist Acantamoeba (32). The functional reconstitution of the highly pure PamNTT1 into artificial lipid vesicles for the first time allowed the biochemical characterization of a representative nonmitochondrial ATP/ADP transporter unaffected by the complex metabolic situation of the bacterial cell. We demonstrated that in contrast to mitochondrial AACs, PamNTT1 catalyzes a membrane potential independent, electroneutral adenine nucleotide hetero-exchange (32, 33). The latter could argue for a cotransport of a counterion compensating for the electrogenic ATP4-/ADP3- exchange.Here, we investigated possible ions accompanying ATP or ADP transport. Interestingly, we uncovered that PamNTT1 and also rickettsial and plastidial ATP/ADP transporters accept an additional important substrate, which is Pi. We performed a comprehensive characterization of the Pi transport and gained new insights into the transport properties of ATP/ADP transporters.  相似文献   

14.
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   

15.
The varitint-waddler mutation A419P renders TRPML3 constitutively active, resulting in cationic overload, particularly in sustained influx of Ca2+. TRPML3 is expressed by inner ear sensory hair cells, and we were intrigued by the fact that hair cells are able to cope with expressing the TRPML3(A419P) isoform for weeks before they ultimately die. We hypothesized that the survival of varitint-waddler hair cells is linked to their ability to deal with Ca2+ loads due to the abundance of plasma membrane calcium ATPases (PMCAs). Here, we show that PMCA2 significantly reduced [Ca2+]i increase and apoptosis in HEK293 cells expressing TRPML3(A419P). The deaf-waddler isoform of PMCA2, operating at 30% efficacy, showed a significantly decreased ability to rescue the Ca2+ loading of cells expressing TRPML3(A419P). When we combined mice heterozygous for the varitint-waddler mutant allele with mice heterozygous for the deaf-waddler mutant allele, we found severe hair bundle defects as well as increased hair cell loss compared with mice heterozygous for each mutant allele alone. Furthermore, 3-week-old double mutant mice lacked auditory brainstem responses, which were present in their respective littermates containing single mutant alleles. Likewise, heterozygous double mutant mice exhibited severe circling behavior, which was not observed in mice heterozygous for TRPML3(A419P) or PMCA2(G283S) alone. Our results provide a molecular rationale for the delayed hair cell loss in varitint-waddler mice. They also show that hair cells are able to survive for weeks with sustained Ca2+ loading, which implies that Ca2+ loading is an unlikely primary cause of hair cell death in ototoxic stress situations.Varitint-waddler (Va) mice express a mutant isoform (A419P) of the transient receptor potential channel TRPML3 (murine gene symbol, Mcoln3) that results in profound hearing loss, vestibular defects (circling behavior, imbalance, head bobbing, waddling), pigmentation deficiencies, sterility, and perinatal lethality in homozygous animals (1). A second Mcoln3 variant (VaJ) that arose in the Va background carries two mutations (I362T and A419P) and shows a phenotype with reduced severity, particularly in heterozygous animals (1). The A419P mutation in Va and VaJ mice is located in transmembrane-spanning domain 5(TM5)3 of TRPML3, where it leads to a constitutively open channel, resulting in highly elevated [Ca2+]i (2-5). In contrast to the effect of the A419P mutation on TRPML3 channel activity, the single I362T mutation does not appear to affect [Ca2+]i (3, 5). When combined with the A419P mutation, as found in VaJ mice, the constitutive activity of this mutant TRPML3 isoform is comparable with that of A419P alone (2-5).Here, we show that HEK293 cells expressing TRPML3-(A419P) or TRPML3(I362T/A419P) undergo rapid apoptosis. This apoptosis is suppressed by coexpression of plasma membrane calcium ATPase type 2 (PMCA2). In varitint-waddler mice, sensory hair cells survive for weeks after birth (6), which raised the question of whether this survival could be the result of the hair cells'' ability to deal with normally transient and localized Ca2+ influx, a feature that is centered around the high levels of mobile Ca2+ buffers and PMCA isoforms found in sensory hair cells (7-10). We decided to test this hypothesis in vivo by utilizing deaf-waddler mice that carry a mutation (G283S) in the Atp2b2 gene encoding mutant PMCA2. Mice homozygous for PMCA2(G283S) (Atp2b2dfw/dfw) are deaf and have poor balance (11). Compared with Atp2b2 knock-out mice, deaf-waddler mice display a milder phenotype because PMCA2(G283S) retains 30% of its biological activity compared with the wild-type isoform (12). We found that sensory hair cell loss, hearing loss, and vestibular dysfunction were aggravated in mice carrying varitint-waddler and deaf-waddler alleles compared with animals carrying the single mutant alleles. Our results reveal that the Ca2+-buffering and Ca2+ extrusion abilities of hair cells are powerful enough to prevent cell death for weeks, even in the presence of constitutively active TRPML3(A419P), which is able to induce rapid apoptosis in other cells.  相似文献   

16.
17.
Voltage dependence and kinetics of CaV1.2 activation are affected by structural changes in pore-lining S6 segments of the α1-subunit. Significant effects are induced by either proline or threonine substitutions in the lower third of segment IIS6 (“bundle crossing region”), where S6 segments are likely to seal the channel in the closed conformation (Hohaus, A., Beyl, S., Kudrnac, M., Berjukow, S., Timin, E. N., Marksteiner, R., Maw, M. A., and Hering, S. (2005) J. Biol. Chem. 280, 38471–38477). Here we report that S435P in IS6 results in a large shift of the activation curve (-25.9 ± 1.2 mV) and slower current kinetics. Threonine substitutions at positions Leu-429 and Leu-434 induced a similar kinetic phenotype with shifted activation curves (L429T by -6.6 ± 1.2 and L434T by -12.1 ± 1.7 mV). Inactivation curves of all mutants were shifted to comparable extents as the activation curves. Interdependence of IS6 and IIS6 mutations was analyzed by means of mutant cycle analysis. Double mutations in segments IS6 and IIS6 induce either additive (L429T/I781T, -34.1 ± 1.4 mV; L434T/I781T, -40.4 ± 1.3 mV; L429T/L779T, -12.6 ± 1.3 mV; and L434T/L779T, -22.4 ± 1.3 mV) or nonadditive shifts of the activation curves along the voltage axis (S435P/I781T, -33.8 ± 1.4 mV). Mutant cycle analysis revealed energetic coupling between residues Ser-435 and Ile-781, whereas other paired mutations in segments IS6 and IIS6 had independent effects on activation gating.Ca2+ current through CaV1.2 channels initiates muscle contraction, release of hormones and neurotransmitters, and affects physiological processes such as vision, hearing, and gene expression (1). Their pore-forming α1-subunit is composed of four homologous domains formed by six transmembrane segments (S1–S6) (2). The signal of the voltage-sensing machinery, consisting of multiple charged amino acids (located in segments S4 and adjacent structures of each domain), is transmitted to the pore region (3). Conformational changes in pore lining S6 and adjacent segments finally lead to pore openings (activation) and closures (inactivation).Our understanding of how CaV1.2 channels open and close is largely based on extrapolations of structural information from potassium channels. The crystal structures of the closed conformation of two bacterial potassium channels (KcsA and MlotiK) (4, 5) show a gate located at the intracellular channel mouth formed by tightly packed S6 helices. The crystal structure of the open conformation of Kv1.2 (6, 7) revealed a bent S6 with the highly conserved PXP motif apparently acting as a hinge (see 8). The activation mechanism proposed for MthK channels involves helix bending at a highly conserved glycine at position 83 (see Ref. 9, “glycine gating hinge” hypothesis).Compared with potassium channels, the pore of CaV is asymmetric, and none of the four S6 segments has a putative helix-bending PXP motif. Furthermore, the conserved glycine (corresponding to position 83 in MthK, see Ref. 10) is only present in segments IS6 and IIS6 (for review see Ref. 11). We have shown that substituting proline for this glycine in IIS6 of CaV1.2 does not significantly affect gating (12).Zhen et al. (13) investigated the pore lining S6 segments of CaV2.1 using the substituted cysteine accessibility method. The accessibility of cysteines was changed by opening and closing the channel, consistent with the gate being on the intracellular side. The general picture of a channel gate close to the inner channel mouth of CaV1.2 was recently supported by pharmacological studies (14).Substitution of hydrophilic residues in the lower third of segment IIS6 of CaV1.2 (LAIA motif, 779–784, see Ref. 12) induces pronounced changes in channel gating as follows: a shift in the voltage dependence of activation accompanied by a slowing of the activation kinetics near the footstep of the m(V) curve and a slowing of deactivation at all potentials. Interestingly, these changes in channel gating resemble the effects of proline substitution of Gly-219 in the bacterial sodium channel from Bacillus halodurans (“Gly-219 gating hinge,” see Ref. 15).The strongest shifts of the activation curve reported so far were observed for proline substitutions (12). As prolines in an α-helix cause a rigid kink with an angle of about 26° (16), we hypothesized that these mutants were causing a kink in helix IIS6 similar to a bend that would normally occur flexibly during the activation process (12).Here we extend our previous study by systematically substituting residues in segment IS6 of CaV1.2 by proline or the small and polar threonine. Several functional IS6 mutants with shifted activation and inactivation characteristics were identified (S435P, L429T, and L434T), and the interdependence of IS6 and IIS6 mutations was analyzed. Mutant cycle analysis revealed both mutually independent and energetically coupled contributions of IS6 and IIS6 residues on activation gating.  相似文献   

18.
The elevation of [cAMP]i is an important mechanism of platelet inhibition and is regulated by the opposing activity of adenylyl cyclase and phosphodiesterase (PDE). In this study, we demonstrate that a variety of platelet agonists, including thrombin, significantly enhance the activity of PDE3A in a phosphorylation-dependent manner. Stimulation of platelets with the PAR-1 agonist SFLLRN resulted in rapid and transient phosphorylation of PDE3A on Ser312, Ser428, Ser438, Ser465, and Ser492, in parallel with the PKC (protein kinase C) substrate, pleckstrin. Furthermore, phosphorylation and activation of PDE3A required the activation of PKC, but not of PI3K/PKB, mTOR/p70S6K, or ERK/RSK. Activation of PKC by phorbol esters also resulted in phosphorylation of the same PDE3A sites in a PKC-dependent, PKB-independent manner. This was further supported by the finding that IGF-1, which strongly activates PI3K/PKB, but not PKC, did not regulate PDE3A. Platelet activation also led to a PKC-dependent association between PDE3A and 14-3-3 proteins. In contrast, cAMP-elevating agents such as PGE1 and forskolin-induced phosphorylation of Ser312 and increased PDE3A activity, but did not stimulate 14-3-3 binding. Finally, complete antagonism of PGE1-evoked cAMP accumulation by thrombin required both Gi and PKC activation. Together, these results demonstrate that platelet activation stimulates PKC-dependent phosphorylation of PDE3A on Ser312, Ser428, Ser438, Ser465, and Ser492 leading to a subsequent increase in cAMP hydrolysis and 14-3-3 binding.Upon vascular injury, platelets adhere to the newly exposed subintimal collagen and undergo activation leading to platelet spreading to cover the damaged region and release of thrombogenic factors such as ADP and thromboxane A2. In addition, platelets are activated by thrombin, which is generated as a result of activation of the coagulation pathway, and stimulates platelets by cleaving the protease-activated receptors (PAR),2 PAR-1 and PAR-4. The final common pathway is the exposure of fibrinogen binding sites on integrin αIIbβ3 resulting in platelet aggregation and thrombus formation.Thrombin-mediated cleavage of PARs leads to activation of phospholipase C β (PLC), hydrolysis of phosphatidylinositol (PI) 4,5-bisphosphate and a subsequent increase in [Ca2+]i and activation of protein kinase C (PKC). Protein kinase C contributes to platelet activation both directly, through affinity regulation of the fibrinogen receptor, integrin αIIbβ3 (1), and indirectly by enhancing degranulation (2). Thrombin also stimulates activation of PI 3-kinases and subsequent generation of PI (3, 4, 5) trisphosphate and PI (3, 4) bisphosphate (3), which recruit protein kinase B (PKB) to the plasma membrane where it becomes phosphorylated and activated.Platelet activation is opposed by agents that raise intracellular 3′-5′-cyclic adenosine monophosphate ([cAMP]i). cAMP is a powerful inhibitory second messenger that down-regulates platelet function by interfering with Ca2+ homeostasis, degranulation and integrin activation (4). Synthesis of cAMP is stimulated by mediators such as prostaglandin I2 (PGI2), which bind to Gs-coupled receptors leading to activation of adenylate cyclase (AC). This inhibitory pathway is opposed by thrombin, which inhibits the elevation of cAMP indirectly via autocrine activation of the Gi-coupled ADP receptor P2Y12. cAMP signaling is terminated by hydrolysis to biologically inert 5′-AMP by 3′-phosphodiesterases. Platelets express two cAMP phosphodiesterase isoforms, cGMP-stimulated PDE2 and cGMP-inhibited PDE3A. PDE3A is the most abundant isoform in platelets and has a ∼250-fold lower Km for cAMP than PDE2 (4). As a consequence of these properties, PDE3A exerts a greater influence on cAMP homeostasis, particularly at resting levels. The importance of PDE3A in platelet function is further emphasized by the finding that the PDE3A inhibitors cilostamide and milrinone raise basal cAMP levels and strongly inhibit thrombin-induced platelet activation (5). Furthermore, PDE3A-/- mice demonstrate increased resting levels of platelet cAMP and are protected against a model of pulmonary thrombosis (6). In contrast, the PDE2 inhibitor EHNA has no significant effect on cAMP levels and platelet aggregation (7, 8). The activity of PDE3A is therefore essential to maintain low equilibrium levels of cAMP and determine the threshold for platelet activation (7).Like its paralogue PDE3B, it has recently become clear that PDE3A activity can be positively regulated by phosphorylation in platelets and human oocytes (9, 10). There is some evidence that PKB may be involved in this regulation, although the phosphorylation sites are poorly characterized. In contrast, phosphorylation of PDE3A in HeLa cells was stimulated by phorbol esters and blocked by inhibitors of PKC (11). In this study, we aimed to identify the signaling pathways and phosphorylation sites that are involved in regulation of platelet PDE3A. Here, we show strong evidence that PKC, and not PKB, is involved in agonist-stimulated PDE3A phosphorylation on Ser312, Ser428, Ser438, Ser465, and Ser492, leading to an increase in PDE3A activity, 14-3-3 binding and modulation of intracellular cAMP levels.  相似文献   

19.
20.
We investigated the role of G protein-coupled receptor kinase (GRK)-mediated phosphorylation in agonist-induced desensitization, arrestin association, endocytosis, and intracellular trafficking of the D2 dopamine receptor (DAR). Agonist activation of D2 DARs results in rapid and sustained receptor phosphorylation that is solely mediated by GRKs. A survey of GRKs revealed that only GRK2 or GRK3 promotes D2 DAR phosphorylation. Mutational analyses resulted in the identification of eight serine/threonine residues within the third cytoplasmic loop of the receptor that are phosphorylated by GRK2/3. Simultaneous mutation of these eight residues results in a receptor construct, GRK(-), that is completely devoid of agonist-promoted GRK-mediated receptor phosphorylation. We found that both wild-type (WT) and GRK(-) receptors underwent a similar degree of agonist-induced desensitization as assessed using [35S]GTPγS binding assays. Similarly, both receptor constructs internalized to the same extent in response to agonist treatment. Furthermore, using bioluminescence resonance energy transfer assays to directly assess receptor association with arrestin3, we found no differences between the WT and GRK(-) receptors. Thus, phosphorylation is not required for arrestin-receptor association or agonist-induced desensitization or internalization. In contrast, when we examined recycling of the D2 DARs to the cell surface, subsequent to agonist-induced endocytosis, the GRK(-) construct exhibited less recycling in comparison with the WT receptor. This impairment appears to be due to a greater propensity of the GRK(-) receptors to down-regulate once internalized. In contrast, if the receptor is highly phosphorylated, then receptor recycling is promoted. These results reveal a novel role for GRK-mediated phosphorylation in regulating the post-endocytic trafficking of a G protein-coupled receptor.Dopamine receptors (DARs)3 are members of the GPCR superfamily and consist of five structurally distinct subtypes (1, 2). These can be divided into two subfamilies on the basis of their structure and pharmacological and transductional properties (3). The “D1-like” subfamily includes the D1 and D5 receptors, which couple to the heterotrimeric G proteins GS or GOLF to stimulate adenylyl cyclase activity and raise intracellular levels of cAMP. The D2-like subfamily includes the D2, D3, and D4 receptors, which couple to inhibitory Gi/o proteins to reduce adenylyl cyclase activity as well as modulate voltage-gated K+ or Ca2+ channels. Within the central nervous system, these receptors modulate movement, learning and memory, reward and addiction, cognition, and certain neurendocrine functions. As with other GPCRs, the DARs are subject to a wide variety of regulatory mechanisms, which can either positively or negatively modulate their expression and functional activity (4).Upon agonist activation, most GPCRs undergo desensitization, a homeostatic process that results in a waning of receptor response despite continued agonist stimulation (5, 6). Desensitization is believed to involve the phosphorylation of receptors by either G protein-coupled receptor kinases (GRKs) and/or second messenger-activated kinases such as PKA or PKC. Homologous forms of desensitization involve only agonist-activated receptors and appear to be primarily mediated by GRKs. In many cases, GRK-mediated phosphorylation has been shown to decrease receptor/G protein interactions and also initiate arrestin binding, which further promotes endocytosis of the receptor through clathrin-coated pits (79). Once internalized, GPCRs can engage additional signaling pathways (10), be sorted for recycling to the plasma membrane, or targeted for degradation (79). Among the DARs, the D2 receptor is arguably one of the most validated drug targets in neurology and psychiatry. For instance, all receptor-based anti-parkinsonian drugs work via stimulating the D2 DAR (11), whereas all Food and Drug Administration-approved antipsychotic agents are antagonists of this receptor subtype (12, 13). The D2 DAR is also therapeutically targeted in other disorders such as restless legs syndrome, tardive dyskinesia, Tourette syndrome, and hyperprolactinemia. As such, more knowledge concerning the regulation of the D2 DAR could be helpful in improving current therapies or devising new treatment strategies.In comparison with other GPCRs, however, detailed mechanistic information concerning regulation of the D2 DAR is mostly lacking, although some progress has recently been made. For instance, we (14) and others (15) have found that PKC-mediated phosphorylation can regulate both D2 receptor desensitization and trafficking. In our PKC study, we mapped out all of the PKC phosphorylation sites within the third intracellular loop (IC3) of the receptor, and we determined the existence of two PKC phosphorylation domains. Both of these domains were found to regulate receptor sequestration, whereas only one domain regulated functional uncoupling in response to PKC activation (14). In response to agonist activation, the D2 DAR has also been shown to undergo functional desensitization (4), although this has not been intensively investigated. More thoroughly examined is the observation that agonist stimulation of the D2 DAR promotes its sequestration from the cell surface into vesicular compartments that appear distinct from those harboring internalized D1 DARs or β-adrenergic receptors (1621). In addition to uncertainty over the endocytic pathway involved, controversy also exists as to whether or not D2 DAR internalization is dynamin-dependent and whether the internalized receptors can partially or completely recycle to the cell surface or, alternatively, if they undergo degradation (19, 2124). The D2 DAR does appear to undergo GRK-mediated phosphorylation upon agonist activation, which has been suggested to promote arrestin association and receptor sequestration (16, 19, 25), although this process has not been studied in detail and its relationship to functional receptor desensitization is unknown.In this study, we have further characterized GRK-mediated phosphorylation of the D2 DAR and determined its role in agonist-induced receptor desensitization, internalization, and recycling. Using site-directed mutagenesis, we have mapped out all of the GRK phosphorylation sites within the D2 DAR and determined that these are distinct from the PKC phosphorylation sites. Using a GRK phosphorylation-null mutant receptor, we found, surprisingly, that GRK-mediated phosphorylation is not actually required for agonist-induced receptor desensitization, arrestin association, or internalization. In contrast, we found that the GRK phosphorylation-null receptor was impaired in its ability to recycle to the cell surface subsequent to internalization and was degraded to a greater extent in comparison with the wild-type receptor. These results suggest that GRK-mediated phosphorylation of the D2 DAR regulates its intracellular trafficking or sorting once internalized, a novel mechanism for GRK-mediated regulation of GPCR function.  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