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 156 毫秒
1.
Neuronal nicotinic acetylcholine receptors (nAChR) composed of α4 + β2 subunits, the high affinity nicotine-binding site in the mammalian brain, up-regulate in response to chronic nicotine exposure. The identities of endogenous mediators of this process are unknown. We find that choline also up-regulates α4 + β2 nAChRs stably expressed by HEK293 cells as measured by increased [3H]epibatidine density. Choline-mediated up-regulation is dose-dependent and corresponds with an increase in β2 subunit protein expression. The choline kinase inhibitor hemicholinium-3 inhibits ∼60% of choline-mediated up-regulation revealing both an HC3-dependent and -independent pathway. Furthermore, choline-mediated up-regulation is not additive with up-regulation agents such as nicotine, but it is additive with weaker promoters of the up-regulation process. When co-applied with the pro-inflammatory cytokine tumor necrosis factor α, choline-mediated up-regulation is increased further through a mechanism that includes an increase in both α4 and β2 protein expression, and this is inhibited by the p38 MAPK inhibitor SB202190. These findings extend the view that up-regulation of α4 + β2 nAChRs is a normal physiological response to altered metabolic and inflammatory conditions.  相似文献   

2.
We employed a pH-sensitive GFP analog, superecliptic phluorin, to observe aspects of nicotinic acetylcholine receptor (nAChR) trafficking to the plasma membrane (PM) in cultured mouse cortical neurons. The experiments exploit differences in the pH among endoplasmic reticulum (ER), trafficking vesicles, and the extracellular solution. The data confirm that few α4β4 nAChRs, but many α4β2 nAChRs, remain in neutral intracellular compartments, mostly the ER. We observed fusion events between nAChR-containing vesicles and PM; these could be quantified in the dendritic processes. We also studied the β4R348C polymorphism, linked to amyotrophic lateral sclerosis (ALS). This mutation depressed fusion rates of α4β4 receptor-containing vesicles with the PM by ∼2-fold, with only a small decrease in the number of nAChRs per vesicle. The mutation also decreased the number of ER exit sites, showing that the reduced receptor insertion results from a change at an early stage in trafficking. We confirm the previous report that the mutation leads to reduced agonist-induced currents; in the cortical neurons studied, the reduction amounts to 2–3-fold. Therefore, the reduced agonist-induced currents are caused by the reduced number of α4β4-containing vesicles reaching the membrane. Chronic nicotine exposure (0.2 μm) did not alter the PM insertion frequency or trafficking behavior of α4β4-laden vesicles. In contrast, chronic nicotine substantially increased the number of α4β2-containing vesicle fusions at the PM; this stage in α4β2 nAChR up-regulation is presumably downstream from increased ER exit. Superecliptic phluorin provides a tool to monitor trafficking dynamics of nAChRs in disease and addiction.  相似文献   

3.
Chronic exposure to nicotine up-regulates high sensitivity nicotinic acetylcholine receptors (nAChRs) in the brain. This up-regulation partially underlies addiction and may also contribute to protection against Parkinson’s disease. nAChRs containing the α6 subunit (α6* nAChRs) are expressed in neurons in several brain regions, but comparatively little is known about the effect of chronic nicotine on these nAChRs. We report here that nicotine up-regulates α6* nAChRs in several mouse brain regions (substantia nigra pars compacta, ventral tegmental area, medial habenula, and superior colliculus) and in neuroblastoma 2a cells. We present evidence that a coat protein complex I (COPI)-mediated process mediates this up-regulation of α6* or α4* nAChRs but does not participate in basal trafficking. We show that α6β2β3 nAChR up-regulation is prevented by mutating a putative COPI-binding motif in the β3 subunit or by inhibiting COPI. Similarly, a COPI-dependent process is required for up-regulation of α4β2 nAChRs by chronic nicotine but not for basal trafficking. Mutation of the putative COPI-binding motif or inhibition of COPI also results in reduced normalized Förster resonance energy transfer between α6β2β3 nAChRs and εCOP subunits. The discovery that nicotine exploits a COPI-dependent process to chaperone high sensitivity nAChRs is novel and suggests that this may be a common mechanism in the up-regulation of nAChRs in response to chronic nicotine.  相似文献   

4.
Nicotine elicits bitter taste by activating TRPM5-dependent and TRPM5-independent but neuronal nAChR-dependent pathways. The nAChRs represent common targets at which acetylcholine, nicotine and ethanol functionally interact in the central nervous system. Here, we investigated if the nAChRs also represent a common pathway through which the bitter taste of nicotine, ethanol and acetylcholine is transduced. To this end, chorda tympani (CT) taste nerve responses were monitored in rats, wild-type mice and TRPM5 knockout (KO) mice following lingual stimulation with nicotine free base, ethanol, and acetylcholine, in the absence and presence of nAChR agonists and antagonists. The nAChR modulators: mecamylamine, dihydro-β-erythroidine, and CP-601932 (a partial agonist of the α3β4* nAChR), inhibited CT responses to nicotine, ethanol, and acetylcholine. CT responses to nicotine and ethanol were also inhibited by topical lingual application of 8-chlorophenylthio (CPT)-cAMP and loading taste cells with [Ca2+]i by topical lingual application of ionomycin + CaCl2. In contrast, CT responses to nicotine were enhanced when TRC [Ca2+]i was reduced by topical lingual application of BAPTA-AM. In patch-clamp experiments, only a subset of isolated rat fungiform taste cells exposed to nicotine responded with an increase in mecamylamine-sensitive inward currents. We conclude that nAChRs expressed in a subset of taste cells serve as common receptors for the detection of the TRPM5-independent bitter taste of nicotine, acetylcholine and ethanol.  相似文献   

5.
Nicotinic acetylcholine receptor (nAChR) agonists, such as epibatidine and its molecular derivatives, are potential therapeutic agents for a variety of neurological disorders. In order to identify determinants for subtype-selective agonist binding, it is important to determine whether an agonist binds in a common orientation in different nAChR subtypes. To compare the mode of binding of epibatidine in a muscle and a neuronal nAChR, we photolabeled Torpedo α2βγδ and expressed human α4β2 nAChRs with [3H]epibatidine and identified by Edman degradation the photolabeled amino acids. Irradiation at 254 nm resulted in photolabeling of αTyr198 in agonist binding site Segment C of the principal (+) face in both α subunits and of γLeu109 and γTyr117 in Segment E of the complementary (−) face, with no labeling detected in the δ subunit. For affinity-purified α4β2 nAChRs, [3H]epibatidine photolabeled α4Tyr195 (equivalent to Torpedo αTyr190) in Segment C as well as β2Val111 and β2Ser113 in Segment E (equivalent to Torpedo γLeu109 and γTyr111, respectively). Consideration of the location of the photolabeled amino acids in homology models of the nAChRs based upon the acetylcholine-binding protein structure and the results of ligand docking simulations suggests that epibatidine binds in a single preferred orientation within the α-γ transmitter binding site, whereas it binds in two distinct orientations in the α4β2 nAChR.Nicotinic acetylcholine receptors (nAChRs)3 are prototypical members of the Cys loop superfamily of neurotransmitter-gated ion channels that mediate the actions of the neurotransmitter acetylcholine (1). nAChRs from vertebrate skeletal muscle and the electric organs of Torpedo rays are heteropentamers of homologous subunits with a stoichiometry of 2α:β:γ(ϵ):δ that are arranged pseudosymmetrically around central cation-selective ion channels (1, 2). There are 12 mammalian neuronal nAChR subunit genes: nine neuronal α subunits (α2–α10) and three neuronal β subunits (β2–β4). The α4β2 nAChR is the most abundant and widely distributed nAChR subtype expressed in the brain and is a major target for potential therapeutic agents for neurological diseases and conditions, including nicotine dependence and Alzheimer and Parkinson diseases (3, 4). Although the ratio of α4 to β2 subunit in vivo is uncertain, expressed receptors containing either three α4 or three β2 subunits have distinct pharmacological properties (5, 6).The agonist binding sites (ABS) of nAChRs are located within the amino-terminal extracellular domain at the interface of adjacent subunits (α-γ and α-δ in the Torpedo nAChR), and different nAChR subunit combinations form ABS with distinct physical and pharmacological properties (3, 7). Affinity labeling studies with Torpedo nAChR and site-directed mutational analyses of muscle and neuronal nAChRs identified key amino acids delineating the ABS from three noncontiguous stretches of the α subunit (Segments A-C, the principal component (+ face)) and three noncontiguous regions of the non-α subunit (Segments D–F, the complementary component (− face)) (8, 9). The three-dimensional structure of the ABS in the absence and presence of nAChR agonists or competitive antagonists has been determined for snail acetylcholine-binding proteins (AChBPs) that are soluble homopentamers homologous to the extracellular (amino-terminal) domain of a nAChR (1012). In the AChBP, four aromatic amino acids from Segments A–C that are conserved within α subunits, along with a conserved Trp in Segment D, form a core aromatic “pocket” with a dimension optimal for accommodation of a trimethylammonium group. The other amino acids in the non-α subunits closest to the aromatic pocket, which are generally not conserved among γ, δ, or neuronal β subunits, are on three antiparallel β strands. The AChBP structure was used to refine the structure of the Torpedo nAChR in the absence of agonist to 4 Å resolution (13). In this structure, there is a reorientation of Segments A–C, resulting in the absence of a well defined core aromatic binding pocket.Analysis of agonist interactions with mutant nAChRs containing fluorine-substituted core aromatic residues provides evidence that cation-π interactions, particularly with αTrp149 in Segment B, are important determinants of agonist binding affinity (14) and for the higher affinity binding of nicotine to α4β2 nAChRs compared with α2βγδ nAChRs (15). Mutational analyses and molecular docking calculations have also provided evidence that two molecules of very similar structure may actually bind to a single receptor in very different orientations, as seen for two high affinity antagonists, d-tubocurarine and its quaternary ammonium analog metocurine, binding to the AChBP and to the muscle nAChR (16, 17).Photoaffinity labeling provides an alternative means to identify amino acids contributing to a drug binding site (18, 19) and has been used to determine the orientation of drugs bound in the ABS of Torpedo nAChR (20). Epibatidine binds with very high affinity (∼10 pm) to heteromeric neuronal nAChRs (e.g. α4β2) and with nanomolar affinity to α7 and muscle-type/Torpedo nAChRs (3). Utilizing a photoreactive analogue of epibatidine (azidoepibatidine; Fig. 1) and mass spectrometry, Tomizawa et al. (21) identified photolabeled amino acids in the Aplysia AChBP (Tyr195 in Segment C and Met116 in Segment E), establishing an orientation for bound azidoepibatidine consistent with the orientation of epibatidine in an AChBP crystal structure (12).Open in a separate windowFIGURE 1.Structure of [3H]epibatidine (top) and azidoepibatidine (bottom).In this report, we use [3H]epibatidine as a photoaffinity reagent to identify the amino acids photolabeled in an expressed α4β2 nAChR and in the Torpedo α2βγδ nAChR. Comparisons of the labeled amino acids seen in the Torpedo nAChR α-γ binding site and in the α4β2 nAChR, in conjunction with the results of docking calculations for epibatidine binding to homology models of the α2βγδ and α4β2 nAChRs, suggests that epibatidine binds in a single orientation in the α-γ site but in two orientations in the α4β2 ABS.  相似文献   

6.
7.
Nicotine increases the number of neuronal nicotinic acetylcholine receptors (nAChRs) in brain. This study investigated the effects of chronic nicotine treatment on nAChRs expressed in primary cultured neurons. In particular, we studied the chronic effects of nicotine exposure on the total density, surface expression and turnover rate of heteromeric nAChRs. The receptor density was measured by [12?I]epibatidine ([12?I]EB) binding. Untreated and nicotine-treated neurons were compared from several regions of embryonic (E19) rat brain. Twelve days of treatment with 10 μM nicotine produced a twofold up-regulation of nAChRs. Biotinylation and whole-cell binding studies indicated that up-regulation resulted from an increase in the number of cell surface receptors as well as intracellular receptors. nAChR subunit composition in cortical and hippocampal neurons was assessed by immunoprecipitation with subunit-selective antibodies. These neurons contain predominantly α4, β2 and α5 subunits, but α2, α3, α6 and β4 subunits were also detected. Chronic nicotine exposure yielded a twofold increase in the β2-containing receptors and a smaller up-regulation in the α4-containing nAChRs. To explore the mechanisms of up-regulation we investigated the effects of nicotine on the receptor turnover rate. We found that the turnover rate of surface receptors was > 2 weeks and chronic nicotine exposure had no effect on this rate.  相似文献   

8.
Nicotine exerts its reinforcing action by stimulating nicotinic acetylcholine receptors (nAChRs) and boosting dopamine (DA) output from the ventral tegmental area (VTA). Recent data have led to a debate about the principal pathway of nicotine action: direct stimulation of the DAergic cells through nAChR activation, or disinhibition mediated through desensitization of nAChRs on GABAergic interneurons. We use a computational model of the VTA circuitry and nAChR function to shed light on this issue. Our model illustrates that the α4β2-containing nAChRs either on DA or GABA cells can mediate the acute effects of nicotine. We account for in vitro as well as in vivo data, and predict the conditions necessary for either direct stimulation or disinhibition to be at the origin of DA activity increases. We propose key experiments to disentangle the contribution of both mechanisms. We show that the rate of endogenous acetylcholine input crucially determines the evoked DA response for both mechanisms. Together our results delineate the mechanisms by which the VTA mediates the acute rewarding properties of nicotine and suggest an acetylcholine dependence hypothesis for nicotine reinforcement.  相似文献   

9.
Human genetic association studies have shown gene variants in the α5 subunit of the neuronal nicotinic receptor (nAChR) influence both ethanol and nicotine dependence. The α5 subunit is an accessory subunit that facilitates α4* nAChRs assembly in vitro. However, it is unknown whether this occurs in the brain, as there are few research tools to adequately address this question. As the α4*-containing nAChRs are highly expressed in the ventral tegmental area (VTA) we assessed the molecular, functional and pharmacological roles of α5 in α4*-containing nAChRs in the VTA. We utilized transgenic mice α5+/+(α4YFP) and α5-/-(α4YFP) that allow the direct visualization and measurement of α4-YFP expression and the effect of the presence (α5+/+) and absence of α5 (-/-) subunit, as the antibodies for detecting the α4* subunits of the nAChR are not specific. We performed voltage clamp electrophysiological experiments to study baseline nicotinic currents in VTA dopaminergic neurons. We show that in the presence of the α5 subunit, the overall expression of α4 subunit is increased significantly by 60% in the VTA. Furthermore, the α5 subunit strengthens baseline nAChR currents, suggesting the increased expression of α4* nAChRs to be likely on the cell surface. While the presence of the α5 subunit blunts the desensitization of nAChRs following nicotine exposure, it does not alter the amount of ethanol potentiation of VTA dopaminergic neurons. Our data demonstrates a major regulatory role for the α5 subunit in both the maintenance of α4*-containing nAChRs expression and in modulating nicotinic currents in VTA dopaminergic neurons. Additionally, the α5α4* nAChR in VTA dopaminergic neurons regulates the effect of nicotine but not ethanol on currents. Together, the data suggest that the α5 subunit is critical for controlling the expression and functional role of a population of α4*-containing nAChRs in the VTA.  相似文献   

10.
11.

Background

Acetamiprid (ACE) and imidacloprid (IMI) belong to a new, widely used class of pesticide, the neonicotinoids. With similar chemical structures to nicotine, neonicotinoids also share agonist activity at nicotinic acetylcholine receptors (nAChRs). Although their toxicities against insects are well established, their precise effects on mammalian nAChRs remain to be elucidated. Because of the importance of nAChRs for mammalian brain function, especially brain development, detailed investigation of the neonicotinoids is needed to protect the health of human children. We aimed to determine the effects of neonicotinoids on the nAChRs of developing mammalian neurons and compare their effects with nicotine, a neurotoxin of brain development.

Methodology/Principal Findings

Primary cultures of cerebellar neurons from neonatal rats allow for examinations of the developmental neurotoxicity of chemicals because the various stages of neurodevelopment—including proliferation, migration, differentiation, and morphological and functional maturation—can be observed in vitro. Using these cultures, an excitatory Ca2+-influx assay was employed as an indicator of neural physiological activity. Significant excitatory Ca2+ influxes were evoked by ACE, IMI, and nicotine at concentrations greater than 1 µM in small neurons in cerebellar cultures that expressed the mRNA of the α3, α4, and α7 nAChR subunits. The firing patterns, proportion of excited neurons, and peak excitatory Ca2+ influxes induced by ACE and IMI showed differences from those induced by nicotine. However, ACE and IMI had greater effects on mammalian neurons than those previously reported in binding assay studies. Furthermore, the effects of the neonicotinoids were significantly inhibited by the nAChR antagonists mecamylamine, α-bungarotoxin, and dihydro-β-erythroidine.

Conclusions/Significance

This study is the first to show that ACE, IMI, and nicotine exert similar excitatory effects on mammalian nAChRs at concentrations greater than 1 µM. Therefore, the neonicotinoids may adversely affect human health, especially the developing brain.  相似文献   

12.
Positive allosteric modulators (PAMs) of nicotinic acetylcholine receptors (nAChR) are important therapeutic candidates as well as valuable research tools. We identified a novel type II PAM, (R)-7-bromo-N-(piperidin-3-yl)benzo[b]thiophene-2-carboxamide (Br-PBTC), which both increases activation and reactivates desensitized nAChRs. This compound increases acetylcholine-evoked responses of α2* and α4* nAChRs but is without effect on α3* or α6* nAChRs (* indicates the presence of other nAChR subunits). Br-BPTC acts from the C-terminal extracellular sequences of α4 subunits, which is also a PAM site for steroid hormone estrogens such as 17β-estradiol. Br-PBTC is much more potent than estrogens. Like 17β-estradiol, the non-steroid Br-PBTC only requires one α4 subunit to potentiate nAChR function, and its potentiation is stronger with more α4 subunits. This feature enables Br-BPTC to potentiate activation of (α4β2)(α6β2)β3 but not (α6β2)2β3 nAChRs. Therefore, this compound is potentially useful in vivo for determining functions of different α6* nAChR subtypes. Besides activation, Br-BPTC affects desensitization of nAChRs induced by sustained exposure to agonists. After minutes of exposure to agonists, Br-PBTC reactivated short term desensitized nAChRs that have at least two α4 subunits but not those with only one. Three α4 subunits were required for Br-BPTC to reactivate long term desensitized nAChRs. These data suggest that higher PAM occupancy promotes channel opening more efficiently and overcomes short and long term desensitization. This C-terminal extracellular domain could be a target for developing subtype or state-selective drugs for nAChRs.  相似文献   

13.
α6β2 Nicotinic acetylcholine receptors (nAChRs) expressed by dopaminergic neurons in the CNS are potential therapeutic targets for the treatment of several neuropsychiatric diseases, including nicotine addiction and Parkinson disease. However, recent studies indicate that the α6 subunit can also associate with the β4 subunit to form α6β4 nAChRs that are difficult to pharmacologically distinguish from α6β2, α3β4, and α3β2 subtypes. The current study characterized a novel 16-amino acid α-conotoxin (α-CTx) TxIB from Conus textile whose sequence is GCCSDPPCRNKHPDLC-amide as deduced from gene cloning. The peptide and an analog with an additional C-terminal glycine were chemically synthesized and tested on rat nAChRs heterologously expressed in Xenopus laevis oocytes. α-CTx TxIB blocked α6/α3β2β3 nAChR with an IC50 of 28 nm. In contrast, the peptide showed little or no block of other tested subtypes at concentrations up to 10 μm. The three-dimensional solution structure of α-CTx TxIB was determined using NMR spectroscopy. α-CTx TxIB represents a uniquely selective ligand for probing the structure and function of α6β2 nAChRs.  相似文献   

14.
Nicotine leads to both activation and desensitization (inactivation) of nicotinic acetylcholine receptors (nAChRs). This study tested the hypothesis that nicotine and a selective antagonist of β2*nAChRs would have similar effects on affective behavior. Adult C57BL/6J male mice were tested in a conditioned emotional response (CER) assay which evaluates the ability of an aversive stimulus to inhibit goal-directed behavior. Mice lever-pressed for a saccharin reinforcer according to a variable schedule of reinforcement during sessions in which two presentations of a compound light/tone conditioned stimulus (CS) co-terminated with a 0.1 or 0.3 mA, 0.5 s footshock unconditioned stimulus (US). During testing in the absence of the US, mice received doses of i.p. nicotine (0, 0.0032, 0.01, 0.032, 0.1 mg/kg) or a selective β2 subunit containing nAChR (β2*nAChR) antagonist dihydro-beta-erythroidine (0, 0.1, 0.3, 1.0, 3.0 mg/kg DHβE). There was a dose-dependent effect of nicotine revealing that only low doses (0.01, 0.032 mg/kg) increased CER suppression ratios (SR) in these mice. DHβE also dose-dependently increased SR at the 3 mg/kg dose. In ethological measures of fear−/anxiety-like behavior, these doses of nicotine and DHβE significantly reduced digging behavior in a marble burying task and 0.3 mg/kg DHβE promoted open-arm activity in the elevated plus maze. Doses of nicotine and DHβE that altered affective behavior had no effect on locomotor activity. Similar to previous reports with anxiolytic drugs, low dose nicotine and DHβE reversed SR in a CER assay, decreased digging in a marble burying assay and increased open arm activity in the elevated plus maze. This study provides evidence that inactivation of β2*nAChRs reduces fear-like and anxiety-like behavior in rodents and suggests that smokers may be motivated to smoke in part to desensitize their β2*nAChRs. These data further identify β2*nAChR antagonism as a potential therapeutic strategy for relief of negative affect and anxiety.  相似文献   

15.
The agonist-binding site of nicotinic acetylcholine receptors (nAChRs) spans an interface between two subunits of the pentameric receptor. The principal component of this binding site is contributed by an α subunit, and it binds the cationic moiety of the nicotinic pharmacophore. The other part of the pharmacophore, a hydrogen bond acceptor, has recently been shown to bind to the complementary non-α subunit via the backbone NH of a conserved Leu. This interaction was predicted by studies of ACh-binding proteins and confirmed by functional studies of the neuronal (CNS) nAChR, α4β2. The ACh-binding protein structures further suggested that the hydrogen bond to the backbone NH is mediated by a water molecule and that a second hydrogen bonding interaction occurs between the water molecule and the backbone CO of a conserved Asn, also on the non-α subunit. Here, we provide new insights into the nature of the interactions between the hydrogen bond acceptor of nicotinic agonists and the complementary subunit backbone. We studied both the nAChR of the neuromuscular junction (muscle-type) and a neuronal subtype, (α4)2(β4)3. In the muscle-type receptor, both ACh and nicotine showed a strong interaction with the Leu NH, but the potent nicotine analog epibatidine did not. This interaction was much attenuated in the α4β4 receptor. Surprisingly, we found no evidence for a functionally significant interaction with the backbone carbonyl of the relevant Asn in either receptor with an array of agonists.  相似文献   

16.
Presynaptic nicotinic acetylcholine receptors (nAChRs) have long been implicated in the modulation of CNS circuits. We previously reported that brief exposure to low concentrations of nicotine induced sustained potentiation of glutamatergic transmission at ventral hippocampal (vHipp)-striatal synapses. Here, we exploited nAChR subtype-selective antagonists and agonists and α7*nAChR knockout mutant mice (α7-/-) to elucidate the signaling mechanisms underlying nAChR-mediated modulation of synaptic transmission. Using a combination of micro-slices culture from WT and α7-/-mice, calcium imaging, and immuno-histochemical techniques, we found that nicotine elicits localized and oscillatory increases in intracellular Ca2+ along vHipp axons that persists for up to 30 minutes. The sustained phase of the nicotine-induced Ca2+ response was blocked by α-BgTx but not by DHβE and was mimicked by α7*nAChR agonists but not by non-α7*nAChR agonists. In vHipp slices from α7-/- mice, nicotine elicited only transient increases of axonal Ca2+ signals and did not activate CaMKII. The sustained phase of the nicotine-induced Ca2+ response required localized activation of CaMKII, phospholipase C, and IP3 receptor mediated Ca2+-induced Ca2+ release (CICR). In conclusion, activation of presynaptic nAChRs by nicotine elicits Ca2+ influx into the presynaptic axons, the sustained phase of the nicotine-induced Ca2+ response requires that axonal α7*nAChR activate a downstream signaling network in the vHipp axons.  相似文献   

17.
We have examined the effect of temperature on the electrophysiological properties of three neuronal nicotinic acetylcholine receptor (nAChR) subtypes: the rapidly desensitizing homomeric α7 nAChR, the more slowly desensitizing heteromeric α4β2 nAChR and on α7 nAChRs containing a transmembrane mutation (L247T) that results in dramatically reduced desensitization. In all cases, the functional properties of receptors expressed in Xenopus oocytes at room temperature (RT; 21°C) were compared to those recorded at either physiological temperature (37°C) or at lower temperature (4°C). Alterations in temperature had dramatically differing effects on the amplitude of whole-cell responses detected with these three nAChR subtypes. Compared to responses at RT, the amplitude of agonist-evoked responses with α4β2 nAChRs was increased at high temperature (125±9%, n = 6, P<0.01) and reduced at low temperature (47±5%, n = 6, P<0.01), whereas the amplitude of α7 responses was reduced at high temperature (27±7%, n = 11, P<0.001) and increased at low temperatures (224±16%, n = 10, P<0.001). In contrast to the effects of temperature on α4β2 and wild type α7 nAChRs, the amplitude of α7 nAChRs containing the L247T mutation was unaffected by changes in temperature. In addition, changes in temperature had little or no effect on current amplitude when α7 nAChRs were activated by the largely non-desensitizing allosteric agonist 4BP-TQS. Despite these differing effects of temperature on the amplitude of agonist-evoked responses in different nAChRs, changes in temperature had a consistent effect on the rate of receptor desensitization on all subtypes examined. In all cases, higher temperature resulted in increased rates of desensitization. Thus, it appears that the differing effects of temperature on the amplitudes of whole-cell responses cannot be explained by temperature-induced changes in receptor desensitization rates.  相似文献   

18.
Epidemiological studies indicate that smoking is negatively correlated with the incidence and development of Alzheimer''s disease (AD). Nicotine was reported to be the active factor. However, the detailed mechanisms still remain to be fully elucidated. Early growth response gene 1 (EGR-1) plays important roles in several important biological processes such as promoting cell growth, differentiation, anti oxidative stress, and apoptosis, but few in the pathogenesis of AD. In the present study, we show that nicotine can activate the MAPK/ERK/EGR-1 signaling pathway partially through α7 nAChR. In addition, the up-regulation of EGR-1 by nicotine can also increase the phosphorylation of CyclinD1 which contributes to the attenuation of amyloid-β (Aβ25–35) -induced neurotoxicity. Although nicotine and Aβ25–35 can activate EGR-1, the expression of EGR-1 is down-regulated following treatment with nicotine and Aβ25–35. This study demonstrates that low dose nicotine attenuates Aβ25–35-induced neurotoxicity in vitro and in vivo through activating EGR-1 pathway.  相似文献   

19.
Nicotinic acetylcholine receptors (nAChRs) expressed in the brain are involved in regulating cognitive functions, as well as inflammatory reactions. Their density is decreased upon Alzheimer disease accompanied by accumulation of β-amyloid (Aβ42), memory deficit and neuroinflammation. Previously we found that α7 nAChR-specific antibody induced pro-inflammatory interleukin-6 production in U373 glioblastoma cells and that such antibodies were present in the blood of humans. We raised a hypothesis that α7 nAChR-specific antibody can cause neuroinflammation when penetrating the brain. To test this, C57Bl/6 mice were either immunized with extracellular domain of α7 nAChR subunit α7(1-208) or injected with bacterial lipopolysaccharide (LPS) for 5 months. We studied their behavior and the presence of α3, α4, α7, β2 and β4 nAChR subunits, Aβ40 and Aβ42 and activated astrocytes in the brain by sandwich ELISA and confocal microscopy. It was found that either LPS injections or immunizations with α7(1-208) resulted in region-specific decrease of α7 and α4β2 and increase of α3β4 nAChRs, accumulation of Aβ42 and activated astrocytes in the brain of mice and worsening of their episodic memory. Intravenously transferred α7 nAChR-specific-antibodies penetrated the brain parenchyma of mice pre-injected with LPS. Our data demonstrate that (1) neuroinflammation is sufficient to provoke the decrease of α7 and α4β2 nAChRs, Aβ42 accumulation and memory impairment in mice and (2) α7(1-208) nAChR-specific antibodies can cause inflammation within the brain resulting in the symptoms typical for Alzheimer disease.  相似文献   

20.
Pathological angiogenesis contributes to tobacco‐related diseases such as malignancy, atherosclerosis and age‐related macular degeneration. Nicotine acts on endothelial nicotinic acetylcholine receptors (nAChRs) to activate endothelial cells and to augment pathological angiogenesis. In the current study, we studied nAChR subunits involved in these actions. We detected mRNA for all mammalian nAChR subunits except α2, α4, γ, and δ in four different types of ECs. Using siRNA methodology, we found that the α7 nAChR plays a dominant role in nicotine‐induced cell signaling (assessed by intracellular calcium and NO imaging, and studies of protein expression and phosphorylation), as well as nicotine‐activated EC functions (proliferation, survival, migration, and tube formation). The α9 and α7 nAChRs have opposing effects on nicotine‐induced cell proliferation and survival. Our studies reveal a critical role for the α7 nAChR in mediating the effects of nicotine on the endothelium. Other subunits play a modulatory role. These findings may have therapeutic implications for diseases characterized by pathological angiogenesis. J. Cell. Biochem. 108: 433–446, 2009. © 2009 Wiley‐Liss, Inc.  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