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.
A comparative denaturation of HbA and HbS in the R states using sodium n-dodecyl sulfate (SDS) was carried out at pH 7.20 in the presence and absence of Calcium (0–40 μM) and monitored by UV–Vis spectrophotometry in the range of 250–650 nm. In the HbS spectra, the calcium alone caused little or no perturbation of the aromatic region but caused a decrease in oxygen affinity when compared to the HbA. The combinations of [SDS] and [Ca] perturbed the HbS the most, relative to the individual spectra of the [SDS] and [Ca]. However, the presence of Ca appeared to diminish the adverse effects of the SDS on HbA. The denaturation pathway of the HbA involved mainly the formation of heme dimers and some ferryl heme species. For the HbS, heme monomers and a large amount of ferryl species were formed. It is suggested that the greater monomer species formed by the HbS denaturation pathway would result in both Fenton and enhanced enzymatic reactions, compared to the dimer. This could lead ultimately to the formation of ferryl radicals. Thus, at physiological pH for the HbS, the Ca–SDS interaction increases the tendency for protein denaturation in comparison to the HbA.  相似文献   

2.
Guanylyl cyclase-activating proteins (GCAPs) are calcium sensor proteins of the EF-hand superfamily that inhibit retinal photoreceptor membrane guanylyl cyclase (retGC) in the dark when they bind Ca(2+) but activate retGC when Ca(2+) dissociates from GCAPs in response to light stimulus. We addressed the difference in exposure of GCAP-2 structure to protein kinase and a protease as indicators of conformational change caused by binding and release of Ca(2+). We have found that unlike its homolog, GCAP-1, the C terminus of GCAP-2 undergoes phosphorylation by cyclic nucleotide-dependent protein kinases (CNDPK) present in the retinal extract and rapid dephosphorylation by the protein phosphatase PP2C present in the retina. Inactivation of the CNDPK phosphorylation site in GCAP-2 by substitutions S201G or S201D, as well as phosphorylation or thiophosphorylation of Ser(201), had little effect on the ability of GCAP-2 to regulate retGC in reconstituted membranes in vitro. At the same time, Ca(2+) strongly inhibited phosphorylation of the wild-type GCAP-2 by retinal CNDPK but did not affect phosphorylation of a constitutively active Ca(2+)-insensitive GCAP-2 mutant. Partial digestion of purified GCAP-2 with Glu-C protease revealed at least two sites that become exposed or constrained in a Ca(2+)-sensitive manner. The Ca(2+)-dependent conformational changes in GCAP-2 affect the areas around Glu(62) residue in the entering helix of EF-hand 2, the areas proximal to the exiting helix of EF-hand 3, and Glu(136)-Glu (138) between EF-hand 3 and EF-hand 4. These changes also cause the release of the C-terminal Ser(201) from the constraint caused by the Ca(2+)-bound conformation.  相似文献   

3.
Alcohol abuse is a leading cause of pancreatitis, accounting for 30% of acute cases and 70-90% of chronic cases, yet the mechanisms leading to alcohol-associated pancreatic injury are unclear. An early and critical feature of pancreatitis is the aberrant signaling of Ca(2+) within the pancreatic acinar cell. An important conductor of this Ca(2+) is the basolaterally localized, intracellular Ca(2+) channel ryanodine receptor (RYR). In this study, we examined the effect of ethanol on mediating both pathologic intra-acinar protease activation, a precursor to pancreatitis, as well as RYR Ca(2+) signals. We hypothesized that ethanol sensitizes the acinar cell to protease activation by modulating RYR Ca(2+). Acinar cells were freshly isolated from rat, pretreated with ethanol, and stimulated with the muscarinic agonist carbachol (1 μM). Ethanol caused a doubling in the carbachol-induced activation of the proteases trypsin and chymotrypsin (p < 0.02). The RYR inhibitor dantrolene abrogated the enhancement of trypsin and chymotrypsin activity by ethanol (p < 0.005 for both proteases). Further, ethanol accelerated the speed of the apical to basolateral Ca(2+) wave from 9 to 18 μm/s (p < 0.0005; n = 18-22 cells/group); an increase in Ca(2+) wave speed was also observed with a change from physiologic concentrations of carbachol (1 μM) to a supraphysiologic concentration (1 mM) that leads to protease activation. Dantrolene abrogated the ethanol-induced acceleration of wave speed (p < 0.05; n = 10-16 cells/group). Our results suggest that the enhancement of pathologic protease activation by ethanol is dependent on the RYR and that a novel mechanism for this enhancement may involve RYR-mediated acceleration of Ca(2+) waves.  相似文献   

4.
This study was conducted to examine the effects of copper on membrane potential and cytosolic free calcium in isolated primary chicken hepatocytes which were exposed to different concentration of Cu(2+) (0, 10, 50, 100 μM) or a mixture of Cu(2+) and vitamin C (50 and 50 μM, respectively). Viability, membrane potential, and cytosolic free Ca(2+) of monolayer cultured hepatocytes were investigated at the indicated time point. Results showed that, among the different concentrations of Cu(2+) exposure, the viability of hepatocytes treated with 100 μM Cu(2+) was the worst at the 12th and 24th hours. The effects of Cu(2+) on viability and proliferation were time and dose dependent. Further investigation indicated that Cu(2+) exposure significantly enhanced cytosolic free Ca(2+) in hepatocytes, compared to that in control group, at the 24th hour. Meanwhile, membrane potential was noticeably reduced in hepatocytes increasing concentration of Cu(2+). Taking these results together, we have shown that Cu(2+) can cause toxicity to primary chicken hepatocytes in excessive dose and the effect of Cu(2+) exposure on membrane potential is not site specific, which is probably mediated by the changes of cytosolic free Ca(2+).  相似文献   

5.
Levetiracetam (LEV), a new anticonvulsant agent primarily used to treat epilepsy, has been used in pain treatment but the cellular mechanism of this action remains unclear. This study aimed to investigate effects of LEV on the excitability and membrane depolarization-induced calcium signaling in isolated rat sensory neurons using the whole-cell patch clamp and fura 2-based ratiometric Ca(2+)-imaging techniques. Dorsal root ganglia (DRG) were excised from neonatal rats, and cultured following enzymatic and mechanical dissociation. Under current clamp conditions, acute application of LEV (30 μM, 100 μM and 300 μM) significantly increased input resistance and caused the membrane to hyperpolarize from resting membrane potential in a dose-dependent manner. Reversal potentials of action potential (AP) after hyperpolarising amplitudes were shifted to more negative, toward to potassium equilibrium potentials, after application of LEV. It also caused a decrease in number of APs in neurons fired multiple APs in response to prolonged depolarization. Fura-2 fluorescence Ca(2+) imaging protocols revealed that HiK(+) (30 mM)-induced intracellular free Ca(2+) ([Ca(2+)](i)) was inhibited to 97.8 ± 4.6% (n = 17), 92.6 ± 4.8% (n = 17, p < 0.01) and 89.1 ± 5.1% (n = 18, p < 0.01) after application of 30 μM, 100 μM and 300 μM LEV (respectively), without any significant effect on basal levels of [Ca(2+)](i). This is the first evidence for the effect of LEV on the excitability of rat sensory neurons through an effect which might involve activation of potassium channels and inhibition of entry of Ca(2+), providing new insights for cellular mechanism(s) of LEV in pain treatment modalities.  相似文献   

6.
Molecular dynamics (MD) simulations have been carried out for 62.5 ps on crystal structures of deoxy sickle cell hemoglobin (HbS) and normal deoxy hemoglobin (HbA) using the CHARMM MD algorithm, with a time step of 0.001 ps. In the trajectory analysis of the 12.5–62.5 (50 ps) simulation, oscillations of the radius of gyration and solvent-accessible surface area were calculated. HbS exhibited a general contraction during the simulation, while HbA exhibited a nearly constant size. The average deviations of simulated structures from the starting structures were found to be 1.8 Å for HbA and 2.3 Å for HbS. The average rms amplitudes of atomic motions (atomic flexibility) were about 0.7 Å for HbA and about 1.0 Å for HbS. The amplitudes of backbone motion correlate well with temperature factors derived from x-ray crystallography. A comparison of flexibility between the α- and β-chains in both HbA and HbS indicates that the β-chains generally exhibited greater flexibility than the α-chains, and that the HbS β-chains exhibit greater flexibility in the N-terminal and D- and F-helix regions than do those of HbA. The average amplitude of backbone torsional oscillations was about 9°, a value comparable with that of other simulations, with enhanced torsional oscillation occurring primarily at the ends of helices or in loop regions between helices. Comparison of atomic flexibility and torsional oscillation results suggests that the increased β-chain flexibility results from relatively concerted motions of secondary structure elements. The increased flexibility may play an important role in HbS polymerization. Time course analysis of conformational energy of association, hydrogen bonding and hydrophobic bonding (as calculated from solvent accessibility) shows that all three of these factors contribute to the stability of subunit association for both hemoglobins. © 1993 John Wiley & Sons, Inc.  相似文献   

7.
Scholten A  Koch KW 《PloS one》2011,6(8):e23117
Zebrafish express in their retina a higher number of guanylate cyclase-activating proteins (zGCAPs) than mammalians pointing to more complex guanylate cyclase signaling systems. All six zGCAP isoforms show distinct and partial overlapping expression profiles in rods and cones. We determined critical Ca(2+)-dependent parameters of their functional properties using purified zGCAPs after heterologous expression in E.coli. Isoforms 1-4 were strong, 5 and 7 were weak activators of membrane bound guanylate cyclase. They further displayed different Ca(2+)-sensitivities of guanylate cyclase activation, which is half maximal either at a free Ca(2+) around 30 nM (zGCAP1, 2 and 3) or around 400 nM (zGCAP4, 5 and 7). Zebrafish GCAP isoforms showed also differences in their Ca(2+)/Mg(2+)-dependent conformational changes and in the Ca(2+)-dependent monomer-dimer equilibrium. Direct Ca(2+)-binding revealed that all zGCAPs bound at least three Ca(2+). The corresponding apparent affinity constants reflect binding of Ca(2+) with high (≤ 100 nM), medium (0.1-5 μM) and/or low (≥ 5 μM) affinity, but were unique for each zGCAP isoform. Our data indicate a Ca(2+)-sensor system in zebrafish rod and cone cells supporting a Ca(2+)-relay model of differential zGCAP operation in these cells.  相似文献   

8.
"Band 3," an integral membrane protein of red blood cells, plays a relevant role in anionic transport. The C- and N-terminal portions of band 3 are cytoplasmatics, and the last is the link site for different glycolitic enzymes, such as glyceraldehyde-3-phosphate dehydrogenase, aldolase, phosphofructokinase, and hemoglobin. All or some of these interactions on the CDB3 protein could allow a subtle modulation of anion flux. The interaction among HbA, Mg(2+), and membrane proteins has been sufficiently investigated, but not the effect of Mg(2+) on pathological hemoglobin in relation to the influx of the SO(4)(2-). The aim of this study was to evaluate the involvement of hemoglobin S in sulfate transport. This has been measured with native and increased concentrations of Mg(2+), using normal erythrocytes containing HbA, sickle red cells containing HbS, or ghosts obtained from both erythrocytes and normal erythrocytes ghosts with HbS added. The magnitude of the SO(4)(2-) rate constant measured in normal red blood cells increased markedly when measured in the presence of varied Mg(2+) concentrations. The results show that a low increase of intracellular Mg(2+) concentrations exercises a different HbA modulation on band 3 protein and consequently higher anion transport activity. The same experiments carried out in sickle red cells showed that the SO(4)(2-) rate constant measured in the presence of native concentrations of Mg(2+) was normal, compared to normal red cells, and was not affected by any increase of intracellular Mg(2+). Our suppositions with regard to the importance exercised by the hemoglobin and the Mg(2+) on the SO(4)(2-) influx were confirmed by comparison of the data obtained through measuring SO(4)(2-) influx with native and increased concentrations of Mg(2+) in both normal and sickle red cell ghosts. Both revealed the same sensitivity to Mg(2+) due to withdrawal of hemoglobins. The incorporation of HbS in normal as well as in sickle red cell ghosts reduced the Mg(2+) response to sulfate influx in both the reconstituted ghosts. Our research demonstrated that the different effects exercised on the rate constants of SO(4)(2-) influx in normal (HbA) and sickle red cells (HbS) by the increased intracellular Mg(2+) could be ascribed to the physical-chemical influence exercised either on the hemoglobins or on the intracellular contents of erythrocytes.  相似文献   

9.
The modulation of the local structure and dynamics of domain III of annexin 2 (Anx2), in both the monomeric (p36) and heterotetrameric forms (p90), by calcium and by membrane binding was studied by time-resolved fluorescence intensity and anisotropy measurements of the single tryptophan residue (W212). The results yield the same dominant excited-state lifetime (1.4 ns) in both p36 and p90, suggesting that the conformation and environment of W212 are very similar. The fluorescence anisotropy decay data were analyzed by associative (two-dimensional) as well as nonassociative (one-dimensional) models. Although no statistical criterion is decisive for one model versus the other, only the associative model allows recovery of a physically relevant value of the Brownian rotational correlation of the protein. Using the associative model, a nanosecond flexibility is detectable in p90 but not in p36. When Ca(2+) binds in the millimolar concentration range to both forms of Anx2, a conformational change takes place leading to an increase of the major excited-state lifetime (2.6 ns) and to a suppression of the W212 local flexibility of p90. Binding to membranes of either p36 or p90 in the presence of Ca(2+) does not induce any conformational change other than that provoked by Ca(2+) binding alone. The W212 local flexibility in both proteins increases significantly, however, in their membrane-bound forms. In the presence of membranes, the conformation change of domain III in p90 displays a sensitivity to Ca(2+) 2 orders of magnitude higher than that of p36, reaching intracellular sub-micromolar concentration ranges. This higher Ca(2+) sensitivity correlates with the Ca(2+)-dependent membrane aggregation but not with their Ca(2+)-dependent binding to membranes. The significance of these structural and dynamical changes for the function of the protein is discussed.  相似文献   

10.
Blebbistatin, a myosin II inhibitor, interferes with myosin-actin interaction and microtubule assembly. By influencing cytoskeletal dynamics blebbistatin counteracts apoptosis of several types of nucleated cells. Even though lacking nuclei and mitochondria, erythrocytes may undergo suicidal cell death or eryptosis, which is characterized by cell shrinkage and cell membrane scrambling with phosphatidylserine exposure at the cell surface. Triggers of eryptosis include energy depletion and osmotic shock, which enhance cytosolic Ca(2+) activity with subsequent Ca(2+)-sensitive cell shrinkage and cell membrane scrambling. The present study explored the effect of blebbistatin on eryptosis. Cell membrane scrambling was estimated from binding of annexin V to phosphatidylserine at the erythrocyte surface, cell volume from forward scatter in fluorescence-activated cell sorting analysis and cytosolic Ca(2+) concentration from Fluo3 fluorescence. Exposure to blebbistatin on its own (1-50 μM) did not significantly modify cytosolic Ca(2+) concentration, forward scatter, or annexin V binding. Glucose depletion (48 h) was followed by a significant increase of Fluo3 fluorescence and annexin V binding, effects significantly blunted by blebbistatin (Fluo3 fluorescence ≥ 25 μM, annexin V binding ≥ 10 μM). Osmotic shock (addition of 550 mM sucrose) again significantly increased Fluo3 fluorescence and annexin binding, effects again significantly blunted by blebbistatin (Fluo3 fluorescence ≥ 25 μM, annexin V binding ≥ 25 μM). The present observations disclose a novel effect of blebbistatin, i.e., an influence on Ca(2+) entry and suicidal erythrocyte death following energy depletion and osmotic shock.  相似文献   

11.
The diheme enzyme MauG catalyzes a six-electron oxidation required for posttranslational modification of a precursor of methylamine dehydrogenase (preMADH) to complete the biosynthesis of its protein-derived tryptophan tryptophylquinone (TTQ) cofactor. The crystal structure of the MauG-preMADH complex revealed the presence of a Ca(2+) in proximity to the two hemes [Jensen, L. M. R., Sanishvili, R., Davidson, V. L., and Wilmot, C. M. (2010) Science 327, 1392-1394]. This Ca(2+) did not readily dissociate; however, after extensive treatment with EGTA or EDTA MauG was no longer able to catalyze TTQ biosynthesis and exhibited altered absorption and resonance Raman spectra. The changes in spectral features are consistent with Ca(2+)-dependent changes in heme spin state and conformation. Addition of H(2)O(2) to the Ca(2+)-depleted MauG did not yield spectral changes characteristic of formation of the bis-Fe(IV) state which is stabilized in native MauG. After addition of Ca(2+) to the Ca(2+)-depleted MauG, full TTQ biosynthesis activity and reactivity toward H(2)O(2) were restored, and the spectral properties returned to those of native MauG. Kinetic and equilibrium studies of Ca(2+) binding to Ca(2+)-depleted MauG indicated a two-step mechanism. Ca(2+) initially reversibly binds to Ca(2+)-depleted MauG (K(d) = 22.4 μM) and is followed by a relatively slow (k = 1.4 × 10(-3) s(-1)) but highly favorable (K(eq) = 4.2) conformational change, yielding an equilibrium dissociation constant K(d,eq) value of 5.3 μM. The circular dichroism spectra of native and Ca(2+)-depleted MauG were essentially the same, consistent with Ca(2+)-induced conformational changes involving domain or loop movements rather than general unfolding or alteration of secondary structure. These results are discussed in the context of the structures of MauG and heme-containing peroxidases.  相似文献   

12.
Hata S  Sorimachi H  Nakagawa K  Maeda T  Abe K  Suzuki K 《FEBS letters》2001,501(2-3):111-114
Calpain, a Ca(2+)-dependent cytosolic cysteine protease, proteolytically modulates specific substrates involved in Ca(2+)-mediated intracellular events, such as signal transduction, cell cycle, differentiation, and apoptosis. The 3D structure of m-calpain, in the absence of Ca(2+), revealed that the two subdomains (domains IIa and IIb) of the protease domain (II) have an 'open' conformation, probably due to interactions with other domains. Although the presence of an EF-hand structure was once predicted in the protease domain, no explicit Ca(2+)-binding structure was identified in the 3D structure. Therefore, it is predicted that if the protease domain is excised from the calpain molecule, it will have a Ca(2+)-independent protease activity. In this study, we have characterized a truncated human m-calpain that consists of only the protease domain. Unexpectedly, the proteolytic activity was Ca(2+)-dependent, very weak, and not effectively inhibited by calpastatin, a calpain inhibitor. Ca(2+)-dependent modification of the protease domain by the cysteine protease inhibitor, E-64c, was clearly observed as a SDS-PAGE migration change, indicating that the conformational changes of this domain are a result of Ca(2+) binding. These results suggest that the Ca(2+) binding to domain II, as well as to domains III, IV, and VI, is critical in the process of complete activation of calpain.  相似文献   

13.
Kostellow AB  Ma GY  Morrill GA 《Steroids》2001,66(11):849-856
Progesterone has been shown to act at plasma membrane receptors on the amphibian oocyte to trigger a cascade of changes in membrane phospholipids and to initiate the G(2)/M transition of the first meiotic division. The earliest event (0-1 min) is the transient N-methylation of phosphatidylethanolamine (PE) to form phosphatidylmonomethylethanolamine (PME), demonstrated using [(3)H]glycerol to prelabel oocyte plasma membrane PE. [(3)H]Glycerol-labeled PME rises 10-fold within the 1-2 min after exposure to progesterone and accounts for conversion of about 50% of the [3H]Glycerol-labeled PE. [(3)H]PME levels slowly decline over the following 10-30 min. [(3)H] or [(14)C] labeled fatty acid experiments showed that newly formed PME is enriched in linoleic or palmitic, but not in arachidonic acid, indicating that specific PE pools undergo progesterone-induced N-methylation. Two plasma membrane changes: activation of serine protease, and Ca(2+) release from the oocyte surface coincide with PME formation; both are prevented by pretreatment of oocytes with the N-methylation inhibitor, 2-methylaminoethane. Media containing PME micelles release both protease and Ca(2+) from intact oocytes within the first 1-2 min. The immediate downstream metabolites of PME, PDE and PC, do not induce serine protease activity or Ca(2+) release. We conclude that progesterone initially activates N-methyltransferase in the oocyte plasma membrane, and that the first product, PME, is responsible for activation of serine protease in the plasma membrane and the release of Ca(2+) from the oocyte surface.  相似文献   

14.
Visinin-like protein 1 (VILIP-1) belongs to the neuronal calcium sensor family of Ca(2+)-myristoyl switch proteins that regulate signal transduction in the brain and retina. Here we analyze Ca(2+) and Mg(2+) binding, characterize metal-induced conformational changes, and determine structural effects of myristoylation and dimerization. Mg(2+) binds functionally to VILIP-1 at EF3 (ΔH = +1.8 kcal/mol and K(D) = 20 μM). Unmyristoylated VILIP-1 binds two Ca(2+) sequentially at EF2 and EF3 (K(EF3) = 0.1 μM and K(EF2) = 1-4 μM), whereas myristoylated VILIP-1 binds two Ca(2+) with lower affinity (K(D) = 1.2 μM) and positive cooperativity (Hill slope = 1.5). NMR assignments and structural analysis indicate that Ca(2+)-free VILIP-1 contains a sequestered myristoyl group like that of recoverin. NMR resonances of the attached myristate exhibit Ca(2+)-dependent chemical shifts and NOE patterns consistent with Ca(2+)-induced extrusion of the myristate. VILIP-1 forms a dimer in solution independent of Ca(2+) and myristoylation. The dimerization site is composed of residues in EF4 and the loop region between EF3 and EF4, confirmed by mutagenesis. We present the structure of the VILIP-1 dimer and a Ca(2+)-myristoyl switch to provide structural insights into Ca(2+)-induced trafficking of nicotinic acetylcholine receptors.  相似文献   

15.
Biliary pancreatitis is the most common etiology for acute pancreatitis, yet its pathophysiological mechanism remains unclear. Ca(2+) signals generated within the pancreatic acinar cell initiate the early phase of pancreatitis, and bile acids can elicit anomalous acinar cell intracellular Ca(2+) release. We previously demonstrated that Ca(2+) released via the intracellular Ca(2+) channel, the ryanodine receptor (RyR), contributes to the aberrant Ca(2+) signal. In this study, we examined whether RyR inhibition protects against pathological Ca(2+) signals, acinar cell injury, and pancreatitis from bile acid exposure. The bile acid tauro-lithocholic acid-3-sulfate (TLCS) induced intracellular Ca(2+) oscillations at 50 μM and a peak-plateau signal at 500 μM, and only the latter induced acinar cell injury, as determined by lactate dehydrogenase (LDH) leakage. Pretreatment with the RyR inhibitors dantrolene or ryanodine converted the peak-plateau signal to a mostly oscillatory pattern (P < 0.05). They also reduced acinar cell LDH leakage, basolateral blebbing, and propidium iodide uptake (P < 0.05). In vivo, a single dose of dantrolene (5 mg/kg), given either 1 h before or 2 h after intraductal TLCS infusion, reduced the severity of pancreatitis down to the level of the control (P < 0.05). These results suggest that the severity of biliary pancreatitis may be ameliorated by the clinical use of RyR inhibitors.  相似文献   

16.
17.
In erythrocytes, spermine concentration decreases gradually with age, which is paralleled by increases of cytosolic Ca2+ concentration, with subsequent cell shrinkage and cell membrane scrambling. Cytosolic Ca2+ was estimated from fluo-3 fluorescence, cell volume from forward scatter, cell membrane scrambling from annexin V binding and cation channel activity with whole-cell patch-clamp in human erythrocytes. Extracellular spermine exerted a dual effect on erythrocyte survival. At 200 μM spermine blunted the increase of intracellular Ca2+, cell shrinkage and annexin V binding following 48 h exposure of cells at +37 °C. In contrast, short exposure (10-30 min) of cells to 2 mM spermine was accompanied by increased cytosolic Ca2+ and annexin binding. Intracellular addition of spermine at subphysiological concentration (0.2 μM) significantly decreased the conductance of monovalent cations (Na+, K+, NMDG+) and of Ca2+. Moreover, spermine (0.2 μM) blunted the stimulation of voltage-independent cation channels by Cl? removal. Spermine (0.2 and 200 μM) added to the extracellular bath solution similarly inhibited the cation conductance in Cl?-containing bath solution. The effect of 0.2 μM spermine, but not the effect of 200 μM, was rapidly reversible. Acute addition (250 μM) of a naphthyl acetyl derivative of spermine (200 μM) again significantly decreased basal cation conductance in NaCl bath solution and inhibited voltage-independent cation channels. Spermine is a powerful regulator of erythrocyte cation channel cytosolic Ca2+ activity and, thus, cell survival.  相似文献   

18.
The thermodynamics of the interaction between Ca(2+) and calmodulin (CaM) was examined using isothermal titration calorimetry (ITC). The chemical denaturation of calmodulin was monitored spectroscopically to determine the stability of Ca(2+)-free (apo) and Ca(2+)-loaded (holo) CaMs. We explored the conformational and structural dynamics of CaM using amide hydrogen-deuterium (H-D) exchange coupled with Fourier transform infrared (FT-IR) spectroscopy. The results of H-D exchange and FT-IR suggest that CaM activation by Ca(2+) binding involves significant conformational changes. The results have also revealed that while the overall conformation of holo-CaM is more stable than that of the apo-CaM, some part of its α-helix structures, most likely the EF-hand domain region, has more solvent exposure, thus, has a faster H-D exchange rate than that of the apo-CaM. The ITC method provides a new strategy for obtaining site-specific Ca(2+) binding properties and a better estimation of the cooperativity and conformational change contributions of coupled EF-hand proteins.  相似文献   

19.
Dicoumarol, a widely used anticoagulant, may cause anemia, which may result from enhanced erythrocyte loss due to bleeding or due to accelerated erythrocyte death. Erythrocytes may undergo suicidal death or eryptosis, characterized by cell shrinkage and phospholipid scrambling of the cell membrane. Eryptosis may be triggered by increase of cytosolic Ca(2+)-activity ([Ca(2+)](i)). The present study explored, whether dicoumarol induces eryptosis. [Ca(2+)](i) was estimated from Fluo3-fluorescence, cation channel activity utilizing whole cell patch clamp, cell volume from forward scatter, phospholipid scrambling from annexin-V-binding, and hemolysis from haemoglobin release. Exposure of erythrocytes for 48 hours to dicoumarol (=10 μM) significantly increased [Ca(2+)](i), enhanced cation channel activity, decreased forward scatter, triggered annexin-V-binding and elicited hemolysis. Following exposure to 30 μM dicoumarol, annexin-V-binding affected approximately 15%, and hemolysis 2% of treated erythrocytes. The stimulation of annexin-V-binding by dicoumarol was abrogated in the nominal absence of Ca(2+). In conclusion, dicoumarol stimulates suicidal death of erythrocytes by stimulating Ca(2+) entry and subsequent triggering of Ca(2+) dependent cell membrane scrambling.  相似文献   

20.
Yang SA  Klee CB 《Biochemistry》2000,39(51):16147-16154
Limited proteolysis of calcineurin in the presence of Ca(2+) suggested that its calmodulin-binding domain, readily degraded by proteases, was unfolded while calcineurin B was compactly folded [Hubbard, M. J., and Klee, C. B. (1989) Biochemistry 28, 1868-1874]. Moreover, in the crystal structure of calcineurin, with the four Ca(2+) sites of calcineurin B occupied, the calmodulin-binding domain is not visible in the electron density map [Kissinger, C. R., et al. (1995) Nature 378, 641-644]. Limited proteolysis of calcineurin in the presence of EGTA, shows that, when the low affinity sites of calcineurin B are not occupied, the calmodulin-binding domain is completely protected against proteolytic attack. Slow cleavages are, however, detected in the linker region between the calmodulin-binding and the autoinhibitory domains of calcineurin A. Upon prolonged exposure to the protease, selective cleavages in carboxyl-terminal end of the first helix and the central helix linker of calcineurin B and the calcineurin B-binding helix of calcineurin A are also detected. Thus, Ca(2+) binding to the low-affinity sites of calcineurin B affects the conformation of calcineurin B and induces a conformational change of the regulatory domain of calcineurin A, resulting in the exposure of the calmodulin-binding domain. This conformational change is needed for the partial activation of the enzyme in the absence of calmodulin and its full activation by calmodulin. A synthetic peptide corresponding to the calmodulin-binding domain is shown to interact with a peptide corresponding to the calcineurin B-binding domain, and this interaction is prevented by calcineurin B in the presence but not the absence of Ca(2+). These observations provide a mechanism to explain the dependence on Ca(2+) binding to calcineurin B for calmodulin activation and for the 10-20-fold increase in affinity of calcineurin for Ca(2+) upon removal of the regulatory domain by limited proteolysis [Stemmer, P. M., and Klee, C. B. (1994) Biochemistry 33, 6859-6866].  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