首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
We investigated self-adhesion between highly negatively charged aggrecan macromolecules extracted from bovine cartilage extracellular matrix by performing atomic force microscopy (AFM) imaging and single-molecule force spectroscopy (SMFS) in saline solutions. By controlling the density of aggrecan molecules on both the gold substrate and the gold-coated tip surface at submonolayer densities, we were able to detect and quantify the Ca2+-dependent homodimeric interaction between individual aggrecan molecules at the single-molecule level. We found a typical nonlinear sawtooth profile in the AFM force-versus-distance curves with a molecular persistence length of lp = 0.31 ± 0.04 nm. This is attributed to the stepwise dissociation of individual glycosaminoglycan (GAG) side chains in aggrecans, which is very similar to the known force fingerprints of other cell adhesion proteoglycan systems. After studying the GAG-GAG dissociation in a dynamic, loading-rate-dependent manner (dynamic SMFS) and analyzing the data according to the stochastic Bell-Evans model for a thermally activated decay of a metastable state under an external force, we estimated for the single glycan interaction a mean lifetime of τ = 7.9 ± 4.9 s and a reaction bond length of xβ = 0.31 ± 0.08 nm. Whereas the xβ-value compares well with values from other cell adhesion carbohydrate recognition motifs in evolutionary distant marine sponge proteoglycans, the rather short GAG interaction lifetime reflects high intermolecular dynamics within aggrecan complexes, which may be relevant for the viscoelastic properties of cartilage tissue.  相似文献   

2.
The kinetics of Ca2+-dependent conformational changes of human cardiac troponin (cTn) were studied on isolated cTn and within the sarcomeric environment of myofibrils. Human cTnC was selectively labeled on cysteine 84 with N-((2-(iodoacetoxy)ethyl)-N-methyl)amino-7-nitrobenz-2-oxa-1,3-diazole and reconstituted with cTnI and cTnT to the cTn complex, which was incorporated into guinea pig cardiac myofibrils. These exchanged myofibrils, or the isolated cTn, were rapidly mixed in a stopped-flow apparatus with different [Ca2+] or the Ca2+-buffer 1,2-Bis(2-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid to determine the kinetics of the switch-on or switch-off, respectively, of cTn. Activation of myofibrils with high [Ca2+] (pCa 4.6) induced a biphasic fluorescence increase with rate constants of >2000 s−1 and ∼330 s−1, respectively. At low [Ca2+] (pCa 6.6), the slower rate was reduced to ∼25 s−1, but was still ∼50-fold higher than the rate constant of Ca2+-induced myofibrillar force development measured in a mechanical setup. Decreasing [Ca2+] from pCa 5.0-7.9 induced a fluorescence decay with a rate constant of 39 s−1, which was approximately fivefold faster than force relaxation. Modeling the data indicates two sequentially coupled conformational changes of cTnC in myofibrils: 1), rapid Ca2+-binding (kB ≈ 120 μM−1 s−1) and dissociation (kD ≈ 550 s−1); and 2), slower switch-on (kon = 390s−1) and switch-off (koff = 36s−1) kinetics. At high [Ca2+], ∼90% of cTnC is switched on. Both switch-on and switch-off kinetics of incorporated cTn were around fourfold faster than those of isolated cTn. In conclusion, the switch kinetics of cTn are sensitively changed by its structural integration in the sarcomere and directly rate-limit neither cardiac myofibrillar contraction nor relaxation.  相似文献   

3.
Dynamic features of Ca2+ interactions with transport and regulatory sites control the Ca2+-fluxes in mammalian Na+/Ca2+(NCX) exchangers bearing the Ca2+-binding regulatory domains on the cytosolic 5L6 loop. The crystal structure of Methanococcus jannaschii NCX (NCX_Mj) may serve as a template for studying ion-transport mechanisms since NCX_Mj does not contain the regulatory domains. The turnover rate of Na+/Ca2+ exchange (kcat = 0.5 ± 0.2s−1) in WT–NCX_Mj is 103–104 times slower than in mammalian NCX. In NCX_Mj, the intrinsic equilibrium (Kint) for bidirectional Ca2+ movements (defined as the ratio between the cytosolic and extracellular Km of Ca2+/Ca2+ exchange) is asymmetric, Kint = 0.15 ± 0.5. Therefore, the Ca2+ movement from the cytosol to the extracellular side is ∼7-times faster than in the opposite direction, thereby representing a stabilization of outward-facing (extracellular) access. This intrinsic asymmetry accounts for observed differences in the cytosolic and extracellulr Km values having a physiological relevance. Bidirectional Ca2+ movements are also asymmetric in mammalian NCX. Thus, the stabilization of the outward-facing access along the transport cycle is a common feature among NCX orthologs despite huge differences in the ion-transport kinetics. Elongation of the cytosolic 5L6 loop in NCX_Mj by 8 or 14 residues accelerates the ion transport rates (kcat) ∼10 fold, while increasing the Kint values 100–250-fold (Kint = 15–35). Therefore, 5L6 controls both the intrinsic equilibrium and rates of bidirectional Ca2+ movements in NCX proteins. Some additional structural elements may shape the kinetic variances among phylogenetically distant NCX variants, although the intrinsic asymmetry (Kint) of bidirectional Ca2+ movements seems to be comparable among evolutionary diverged NCX variants.  相似文献   

4.
5.
Although many synthetic calcium indicators are available, a search for compounds with improved characteristics continues. Here, we describe the synthesis and properties of Asante Calcium Red-1 (ACR-1) and its low affinity derivative (ACR-1-LA) created by linking BAPTA to seminaphthofluorescein. The indicators combine a visible light (450–540 nm) excitation with deep-red fluorescence (640 nm). Upon Ca2+ binding, the indicators raise their fluorescence with longer excitation wavelengths producing higher responses. Although the changes occur without any spectral shifts, it is possible to ratio Ca2+-dependent (640 nm) and quasi-independent (530 nm) emission when using visible (<490 nm) or multiphoton (∼780 nm) excitation. Therefore, both probes can be used as single wavelength or, less dynamic, ratiometric indicators. Long indicator emission might allow easy [Ca2+]i measurement in GFP expressing cells. The indicators bind Ca2+ with either high (Kd = 0.49 ± 0.07 μM; ACR-1) or low affinity (Kd = 6.65 ± 0.13 μM; ACR-1-LA). Chelating Zn2+ (Kd = 0.38 ± 0.02 nM) or Mg2+ (Kd ∼ 5 mM) slightly raises and binding Co2+ quenches dye fluorescence. New indicators are somewhat pH-sensitive (pKa = 6.31 ± 0.07), but fairly resistant to bleaching. The probes are rather dim, which combined with low AM ester loading efficiency, might complicate in situ imaging. Despite potential drawbacks, ACR-1 and ACR-1-LA are promising new calcium indicators.  相似文献   

6.
Connexin hemichannels are regulated by several gating mechanisms, some of which depend critically on the extracellular Ca2+ concentration ([Ca2+]e). It is well established that hemichannel activity is inhibited at normal (∼1 mM) [Ca2+]e, whereas lowering [Ca2+]e to micromolar levels fosters hemichannel opening. Atomic force microscopy imaging shows significant and reversible changes of pore diameter at the extracellular mouth of Cx26 hemichannels exposed to different [Ca2+]e, however, the underlying molecular mechanisms are not fully elucidated. Analysis of the crystal structure of connexin 26 (Cx26) gap junction channels, corroborated by molecular dynamics (MD) simulations, suggests that several negatively charged amino acids create a favorable environment for low-affinity Ca2+ binding within the extracellular vestibule of the Cx26 hemichannel. In particular a highly conserved glutammic acid, found in position 47 in most connexins, is thought to undergo post translational gamma carboxylation (γGlu47), and is thus likely to play an important role in Ca2+ coordination. γGlu47 may also form salt bridges with two conserved arginines (Arg75 and Arg184 in Cx26), which are considered important in stabilizing the structure of the extracellular region.  相似文献   

7.
Solutions of intact cardiac thin filaments were examined with transmission electron microscopy, dynamic light scattering (DLS), and particle-tracking microrheology. The filaments self-assembled in solution with a bell-shaped distribution of contour lengths that contained a population of filaments of much greater length than the in vivo sarcomere size (∼1 μm) due to a one-dimensional annealing process. Dynamic semiflexible modes were found in DLS measurements at fast timescales (12.5 ns-0.0001 s). The bending modulus of the fibers is found to be in the range 4.5-16 × 10−27 Jm and is weakly dependent on calcium concentration (with Ca2+ ≥ without Ca2+). Good quantitative agreement was found for the values of the fiber diameter calculated from transmission electron microscopy and from the initial decay of DLS correlation functions: 9.9 nm and 9.7 nm with and without Ca2+, respectively. In contrast, at slower timescales and high polymer concentrations, microrheology indicates that the cardiac filaments act as short rods in solution according to the predictions of the Doi-Edwards chopsticks model (viscosity, η ∼ c3, where c is the polymer concentration). This differs from the semiflexible behavior of long synthetic actin filaments at comparable polymer concentrations and timescales (elastic shear modulus, G′ ∼ c1.4, tightly entangled) and is due to the relative ratio of the contour lengths (∼30). The scaling dependence of the elastic shear modulus on the frequency (ω) for cardiac thin filaments is G′ ∼ ω3/4 ± 0.03, which is thought to arise from flexural modes of the filaments.  相似文献   

8.
Classic calcium hypothesis states that depolarization-induced increase in intracellular Ca2+ concentration ([Ca2+]i) triggers vesicle exocytosis by increasing vesicle release probability in neurons and neuroendocrine cells. The extracellular Ca2+, in this calcium hypothesis, serves as a reservoir of Ca2+ source. Recently we find that extracellular Ca2+per se inhibits the [Ca2+]i dependent vesicle exocytosis, but it remains unclear whether quantal size is regulated by extracellular, or intracellular Ca2+ or both [1]. In this work we showed that, in physiological condition, extracellular Ca2+per se specifically inhibited the quantal size of single vesicle release in rat adrenal slice chromaffin cells. The extracellular Ca2+ in physiological concentration (2.5 mM) directly regulated fusion pore kinetics of spontaneous quantal release of catecholamine. In addition, removal of extracellular Ca2+ directly triggered vesicle exocytosis without eliciting intracellular Ca2+. We propose that intracellular Ca2+ and extracellular Ca2+per se cooperately regulate single vesicle exocytosis. The vesicle release probability was jointly modulated by both intracellular and extracellular Ca2+, while the vesicle quantal size was mainly determined by extracellular Ca2+ in chromaffin cells physiologically.  相似文献   

9.
Leung KW  Leung FP  Huang Y  Mak NK  Wong RN 《FEBS letters》2007,581(13):2423-2428
We demonstrated that ginsenoside-Re (Re), a pharmacological active component of ginseng, is a functional ligand of glucocorticoid receptor (GR) using competitive ligand-binding assay (IC50 = 156.6 nM; Kd = 49.7 nM) and reporter gene assay. Treatment with Re (1 μM) raises intracellular Ca2+ ([Ca2+]i) and nitric oxide (NO) levels in human umbilical vein endothelial cells as measured using fura-2 and 4-amino-5-methylamino-2′,7′-difluorofluorescein diacetate, respectively. Western blot analysis shows that Re increased phosphorylation of endothelial nitric oxide synthase. These effects were abolished by GR antagonist RU486, siRNA targeting GR, non-selective cation channel blocker 2-aminoethyldiphenylborate, or in the absence of extracellular Ca2+, indicating Re is indeed an agonistic ligand for the GR and the activated GR induces rapid Ca2+ influx and NO production in endothelial cells.  相似文献   

10.
Voltage-gated Ca2+ channels (VGCCs) are recognized for their superb ability for the preferred passage of Ca2+ over any other more abundant cation present in the physiological saline. Most of our knowledge about the mechanisms of selective Ca2+ permeation through VGCCs was derived from the studies on native and recombinant L-type representatives. However, the specifics of the selectivity and permeation of known recombinant T-type Ca2+-channel α1 subunits, Cav3.1, Cav3.2 and Cav3.3, are still poorly defined. In the present study we provide comparative analysis of the selectivity and permeation Cav3.1, Cav3.2, and Cav3.3 functionally expressed in Xenopus oocytes. Our data show that all Cav3 channels select Ca2+ over Na+ by affinity. Cav3.1 and Cav3.2 discriminate Ca2+, Sr2+ and Ba2+ based on the ion's effects on the open channel probability, whilst Cav3.3 discriminates based on the ion's intrapore binding affinity. All Cav3s were characterized by much smaller difference in the KD values for Na+ current blockade by Ca2+ (KD1 ∼ 6 μM) and for Ca2+ current saturation (KD2 ∼ 2 mM) as compared to L-type channels. This enabled them to carry notable mixed Na+/Ca2+ current at close to physiological Ca2+ concentrations, which was the strongest for Cav3.3, smaller for Cav3.2 and the smallest for Cav3.1. In addition to intrapore Ca2+ binding site(s) Cav3.2, but not Cav3.1 and Cav3.3, is likely to possess an extracellular Ca2+ binding site that controls channel permeation. Our results provide novel functional tests for identifying subunits responsible for T-type Ca2+ current in native cells.  相似文献   

11.
Cohesive gels have been obtained by de-esterification of 1.0 wt % high-methoxy citrus pectin (degree of esterification ≈ 68%) in the presence of Ca2+ cations, using a commercial preparation (NovoShape) of fungal methyl esterase cloned from Aspergillus aculeatus. A convenient rate of network formation (gelation within ∼30 min) was achieved at an enzyme concentration of 0.2 PEU/g pectin. At a Ca2+-concentration of 40 mM and incubation temperature of 20 °C, severe syneresis (>7% of sample mass) was observed, but release of fluid decreased with decreasing concentration of Ca2+ and increasing temperature of incubation, becoming undetectable for 10 mM Ca2+ at 30 °C. Under these conditions, progressive development of solid-like character (storage modulus, G′) was observed during 160 min of enzymic de-esterification, and the mechanical spectrum recorded at the end of the incubation period had the form typical of a biopolymer gel. On subsequent heating to 70 °C, dissociation of the gel network (sigmoidal reduction in G′ and G″) was observed. At or above the midpoint temperature of this melting process (∼50 °C), there was no indication of gel formation on enzymic de-esterification (at 50 or 60 °C). At lower temperatures (20, 30 and 40 °C), the rate of gelation (assessed visually) showed no systematic increase as the incubation temperature was increased towards the temperature-optimum of the enzyme (∼50 °C). This unexpected behaviour is attributed to competition between faster de-esterification and slower formation of Ca2+-induced ‘egg-box’ junctions.  相似文献   

12.
The structure of aggregates formed due to DNA interaction with dioleoylphosphatidylcholine (DOPC) vesicles in presence of Ca2+ and Mg2+ cations was investigated using synchrotron small-angle X-ray diffraction. For DOPC/DNA = 1:1 mol/base and in the range of concentration of the cation2+ 0-76.5 mM, the diffractograms show the coexistence of two lamellar phases: Lx phase with repeat distance dLx ∼ 8.26-7.39 nm identified as a phase where the DNA strands are intercalated in water layers between adjacent lipid bilayers, and LDOPC phase with repeat distance dDOPC ∼ 6.45-5.65 nm identified as a phase of partially dehydrated DOPC bilayers without any divalent cations and DNA strands. The coexistence of these phases was investigated as a function of DOPC/DNA molar ratio, length of DNA fragments and temperature. If the amount of lipid increases, the fraction of partially dehydrated LDOPC phase is limited, depends on the portion of DNA in the sample and also on the length of DNA fragments. Thermal behaviour of DOPC + DNA + Ca2+ aggregates was investigated in the range 20-80 °C. The transversal thermal expansivities of both phases were evaluated.  相似文献   

13.
Sergio de la Fuente 《BBA》2010,1797(10):1727-1735
We have investigated the kinetics of mitochondrial Ca2+ influx and efflux and their dependence on cytosolic [Ca2+] and [Na+] using low-Ca2+-affinity aequorin. The rate of Ca2+ release from mitochondria increased linearly with mitochondrial [Ca2+] ([Ca2+]M). Na+-dependent Ca2+ release was predominant al low [Ca2+]M but saturated at [Ca2+]M around 400 μM, while Na+-independent Ca2+ release was very slow at [Ca2+]M below 200 μM, and then increased at higher [Ca2+]M, perhaps through the opening of a new pathway. Half-maximal activation of Na+-dependent Ca2+ release occurred at 5-10 mM [Na+], within the physiological range of cytosolic [Na+]. Ca2+ entry rates were comparable in size to Ca2+ exit rates at cytosolic [Ca2+] ([Ca2+]c) below 7 μM, but the rate of uptake was dramatically accelerated at higher [Ca2+]c. As a consequence, the presence of [Na+] considerably reduced the rate of [Ca2+]M increase at [Ca2+]c below 7 μM, but its effect was hardly appreciable at 10 μM [Ca2+]c. Exit rates were more dependent on the temperature than uptake rates, thus making the [Ca2+]M transients to be much more prolonged at lower temperature. Our kinetic data suggest that mitochondria have little high affinity Ca2+ buffering, and comparison of our results with data on total mitochondrial Ca2+ fluxes indicate that the mitochondrial Ca2+ bound/Ca2+ free ratio is around 10- to 100-fold for most of the observed [Ca2+]M range and suggest that massive phosphate precipitation can only occur when [Ca2+]M reaches the millimolar range.  相似文献   

14.
The PPARγ agonist Rosiglitazone exerts anti-hyperglycaemic effects by regulating the long-term expression of genes involved in metabolism, differentiation and inflammation. In the present study, Rosiglitazone treatment rapidly inhibited (5-30 min) the ER Ca2+ ATPase SERCA2b in monocytic cells (IC50 = 1.88 μM; p < 0.05), thereby disrupting short-term Ca2+ homeostasis (resting [Ca2+]cyto = 121.2 ± 2.9% basal within 1 h; p < 0.05). However, extended Rosiglitazone treatment (72 h) induced dose-dependent SERCA2b up-regulation, and restored calcium homeostasis, in monocytic cells (SERCA2b mRNA: 138.7 ± 5.7% basal (1 μM)/215.0 ± 30.9% basal (10 μM); resting [Ca2+]cyto = 97.3 ± 8.3% basal (10 μM)). As unfavourable cardiovascular outcomes, possibly related to disrupted cellular Ca2+ homeostasis, have been linked to Rosiglitazone, this effect may be of clinical interest. In contrast, in PPRE-luciferase reporter-gene assays, Rosiglitazone induced non-dose-dependent PPARγ-dependent effects (1 μM: 152.5 ± 4.9% basal; 10 μM: 136.1 ± 5.1% basal (p < 0.05 for 1 μM vs. 10 μM)). Thus, we conclude that Rosiglitazone can exert PPARγ-independent non-genomic effects, such as the SERCA2b inhibition seen here, but that long-term Rosiglitazone treatment did not perturb resting [Ca]cyto in this study.  相似文献   

15.
A current popular model to explain phosphorylation of smooth muscle myosin (SMM) by myosin light-chain kinase (MLCK) proposes that MLCK is bound tightly to actin but weakly to SMM. We found that MLCK and calmodulin (CaM) co-purify with unphosphorylated SMM from chicken gizzard, suggesting that they are tightly bound. Although the MLCK:SMM molar ratio in SMM preparations was well below stoichiometric (1:73 ± 9), the ratio was ∼ 23-37% of that in gizzard tissue. Fifteen to 30% of MLCK was associated with CaM at ∼ 1 nM free [Ca2+]. There were two MLCK pools that bound unphosphorylated SMM with Kd ∼ 10 and 0.2 μM and phosphorylated SMM with Kd ∼ 20 and 0.2 μM. Using an in vitro motility assay to measure actin sliding velocities, we showed that the co-purifying MLCK-CaM was activated by Ca2+ and phosphorylation of SMM occurred at a pCa50 of 6.1 and at a Hill coefficient of 0.9. Similar properties were observed from reconstituted MLCK-CaM-SMM. Using motility assays, co-sedimentation assays, and on-coverslip enzyme-linked immunosorbent assays to quantify proteins on the motility assay coverslip, we provide strong evidence that most of the MLCK is bound directly to SMM through the telokin domain and some may also be bound to both SMM and to co-purifying actin through the N-terminal actin-binding domain. These results suggest that this MLCK may play a role in the initiation of contraction.  相似文献   

16.
Rewarming patients from accidental hypothermia are regularly complicated with cardiovascular instability ranging from minor depression of cardiac output to fatal circulatory collapse also termed “rewarming shock”. Since altered Ca2+ handling may play a role in hypothermia-induced heart failure, we studied changes in Ca2+ homeostasis in in situ hearts following hypothermia and rewarming. A rat model designed for studies of the intact heart in a non-arrested state during hypothermia and rewarming was used. Rats were core cooled to 15 °C, maintained at 15 °C for 4 h and thereafter rewarmed. As time-matched controls, one group of animals was kept at 37 °C for 5 h. Total intracellular myocardial Ca2+ content ([Ca2+]i) was measured using 45Ca2+. Following rewarming we found a significant reduction of stroke volume and cardiac output compared to prehypothermic control values as well as to time-matched controls. Likewise, we found that hypothermia and rewarming resulted in a more than six-fold increase in [Ca2+]i to 3.01 ± 0.43 μmol/g dry weight compared to 0.44 ± 0.05 μmol/g dry weight in normothemia control. These findings indicate that hypothermia-induced alterations in the Ca2+-handling result in Ca2+ overload during hypothermia, which may contribute to myocardial failure during and after rewarming.  相似文献   

17.
The HCO3 anion activates sperm motility, an important early step in capacitation, by increasing flagellar beat frequency through a pathway that requires the atypical adenylyl cyclase SACY and the sperm-specific Cα2 catalytic subunit of PKA. Here we show that the accelerating action of HCO3 also requires the continued presence of external Ca2+ (EC50 ∼ 0.5 mM), and find that Ca2+ can be replaced by Sr2+ but not by Mn2+. Ca2+ is required for HCO3 to elevate cAMP, but not for cAMP-AM to increase beat frequency, indicating that external Ca2+ acts before rather than after stimulation of SACY by HCO3. With external Ca2+ present, HCO3 does not alter cytosolic or near-membrane [Ca2+]. Removal of external Ca2+ initiates a slow decline in intracellular [Ca2+] and rapid block of the HCO3-evoked acceleration that is not relieved upon increasing internal [Ca2+] by rapid photolysis of caged Ca2+. We also find that the rapid (t1/2 ∼ 10 s) accelerating action of HCO3 is slowed more than three-fold by the carbonic anhydrase inhibitor acetazolamide. It is unaltered by the broad spectrum anion transport inhibitor SITS, and is not accompanied by detectable changes in intracellular pH. We propose that external Ca2+ binds an unidentified extracellular protein that is required for HCO3 to engage cAMP-mediated activation of motility.  相似文献   

18.
Pseudomonas aeruginosa is an opportunistic human pathogen causing severe acute and chronic infections. Earlier we have shown that calcium (Ca2+) induces P. aeruginosa biofilm formation and production of virulence factors. To enable further studies of the regulatory role of Ca2+, we characterized Ca2+ homeostasis in P. aeruginosa PAO1 cells. By using Ca2+-binding photoprotein aequorin, we determined that the concentration of free intracellular Ca2+ ([Ca2+]in) is 0.14 ± 0.05 μM. In response to external Ca2+, the [Ca2+]in quickly increased at least 13-fold followed by a multi-phase decline by up to 73%. Growth at elevated Ca2+ modulated this response. Treatment with inhibitors known to affect Ca2+ channels, monovalent cations gradient, or P-type and F-type ATPases impaired [Ca2+]in response, suggesting the importance of the corresponding mechanisms in Ca2+ homeostasis. To identify Ca2+ transporters maintaining this homeostasis, bioinformatic and LC–MS/MS-based membrane proteomic analyses were used. [Ca2+]in homeostasis was monitored for seven Ca2+-affected and eleven bioinformatically predicted transporters by using transposon insertion mutants. Disruption of P-type ATPases PA2435, PA3920, and ion exchanger PA2092 significantly impaired Ca2+ homeostasis. The lack of PA3920 and vanadate treatment abolished Ca2+-induced swarming, suggesting the role of the P-type ATPase in regulating P. aeruginosa response to Ca2+.  相似文献   

19.
Li B  Dong L  Fu H  Wang B  Hertz L  Peng L 《Cell calcium》2011,50(1):42-53
Primary cultures of mouse astrocytes were used to investigate effects by chronic treatment (3-21 days) with fluoxetine (0.5-10 μM) on capacitative Ca2+ influx after treatment with the SERCA inhibitor thapsigargin and on receptor agonist-induced increases in free cytosolic Ca2+ concentration [Ca2+]i, determined with Fura-2. The agonists were the 5-HT2B agonist fluoxetine, the α2-adrenergic agonist dexmedetomidine, and ryanodine receptor (RyR) and IP3 receptor (IP3R) agonists. In untreated sister cultures each agonist distinctly increased [Ca2+]i, but in cultures treated for sufficient length of time or with sufficiently high doses of fluoxetine, acute administration of fluoxetine, dexmedetomidine, or RyR or IP3R agonists elicited reduced, in some cases abolished, effects. Capacitative Ca2+ entry, meditated by TRPC1 channels, was sufficiently inhibited to cause a depletion of Ca2+ stores, which could explain the reduced agonist effects. All effects of chronic fluoxetine administration could be replicated by TRPC1 channel antibody or siRNA. Since increases in astrocytic [Ca2+]i regulate release of gliotransmitters, these effects may have profound effects on brain function. They may be important for therapeutic effects of all 5 conventional ‘serotonin-specific reuptake inhibitors’ (SSRIs), which at concentrations used therapeutically (∼1 μM) share other of fluoxetine's chronic effects (Zhang et al., Neuron Glia Biol. 16 (2010) 1-13).  相似文献   

20.
Human EFHC1 is a member of the EF-hand superfamily of Ca2+-binding proteins with three DM10 domains of unclear function. Point mutations in the EFHC1 gene are related to juvenile myoclonic epilepsy, a fairly common idiopathic generalized epilepsy. Here, we report the first structural and thermodynamic analyses of the EFHC1C-terminus (residues 403-640; named EFHC1C), comprising the last DM10 domain and the EF-hand motif. Circular dichroism spectroscopy revealed that the secondary structure of EFHC1C is composed by 34% of α-helices and 17% of β-strands. Size exclusion chromatography and mass spectrometry showed that under oxidizing condition EFHC1C dimerizes through the formation of disulfide bond. Tandem mass spectrometry (MS/MS) analysis of peptides generated by trypsin digestion suggests that the Cys575 is involved in intermolecular S-S bond. In addition, DTNB assay showed that each reduced EFHC1C molecule has one accessible free thiol. Isothermal titration calorimetry (ITC) showed that while the interaction between Ca2+ and EFHC1C is enthalpically driven (ΔH = −58.6 to −67 kJ/mol and TΔS = −22.5 to −31 kJ/mol) the interaction between Mg2+ and EFHC1C involves an entropic gain, and is ∼5 times less enthalpically favorable (ΔH = −11.7 to −14 kJ/mol and TΔS = 21.9 to 19 kJ/mol) than for Ca2+ binding. It was also found that under reducing condition Ca2+ or Mg2+ ions bind to EFHC1C in a 1/1 molar ratio, while under oxidizing condition this ratio is reduced, showing that EFHC1C dimerization blocks Ca2+ and Mg2+ binding.  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