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.
Different (iso)guanosine-based self-assembled ionophores give distinctly different results in extraction experiments with alkali(ne earth) cations. A lipophilic guanosine derivative gives good extraction results for K+, Rb+, Ca2+, Sr2+, and Ba2+ and in competition experiments it clearly favors the divalent Sr2+ (and Ba2+) cations. 1,3-Alternate calix[4]arene tetraguanosine hardly shows any improvement in the extraction percentages compared to its reference compound 1,3-alternate calix[4]arene tetraamide. This indicates that one G-quartet does not provide efficient cation complexation under these conditions. In the case of the lipophilic isoguanosine derivative there is a cation size dependent affinity for the monovalent cations (Cs+ ? Rb+ ? K+), but not for the divalent cations (Ca2+ > Ba2+ > Sr2+ > Mg2+). In competition experiments the isoguanosine derivative, unlike guanosine, does not discriminate between monovalent and divalent cations, giving an almost equal extraction of Cs+ and Ba2+.  相似文献   

2.
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   

3.
Many of the structural domains involved in Ca2+ channel (CACN) inactivation are also involved in determining their sensitivity to antagonist inhibition. We hypothesize that differences in inactivation properties and their structural determinants may suggest candidate domains as targets for the development of novel, selective antagonists. The characteristics of Ca2+ current (ICa) inactivation, steady-state inactivation (SSIN), and recovery from inactivation were studied in freshly dispersed smooth muscle cells from rabbit portal vein (RPV) using whole-cell, voltage-clamp methods. The time course of inactivation could be represented by two time constants. Increasing ICa by increasing [Ca2+]o or with more negative holding potentials decreased both time constants. With Sr2+, Ba2+, or Na+ as the charge carrier, ICa inactivation was also represented by two time constants, both of which were larger than those found with Ca2+. With Ca2+, Sr2+, or Ba2+ as the charge carrier, both time constants had minimum values near the voltage associated with maximum current. When Na+ (140 mM) was the charge carrier, voltages for Imax (−20 mV) or τmin (o mV) did not correspond. SSIN of ICa had a half-maximum voltage of −32±4 mV for Ca2+, −43 mV±5 mV for Sr2+, −41±5 mV for Ba2+, and −68±6 mV for Na+. The slope factor for SSIN per e-fold voltage change was 6.5±0.2 mV for Ca2+, 6.8±0.3 for Sr2+, and 6.6±0.2 for Ba2+, representing four equivalent charges. When Na+ or Li+ was the charge carrier, the slope factor was 13.5±0.7 mV, representing two equivalent charges. For ICa in rat left ventricular (rLV) myocytes, there was no difference in the slope factor of SSIN for Ca2+ and Na+. The rate of recovery of ICa from inactivation varied inversely with recovery voltage and was independent of the charge carrier. These results suggest that inactivation of ICa in PV myocytes possess an intrinsic voltage dependence that is modified by Ca2+. For RPV but not rLV ICa, the charge of the permeating ion confers the voltage-dependency of SSIN.  相似文献   

4.
Despite the progress in studies of the properties and functions of low-threshold calcium channels (LTCCs) [1], the mechanisms of their selectivity and permeability remain unstudied in detail. We performed a comparative analysis of the selectivity of three cloned pore-forming LTCC subunits (α1G, α1H, and α1I) functionally expressed in Xenopus oocytes with respect to bivalent alkaline-earth metal cations (Ba2+, Ca2+, and Sr2+. The relative conductivities (G) of these channels were determined according to the amplitudes of macroscopic currents (I) and potentials of zero currents (E). The currents were recorded after preliminary intracellular injection of a fast calcium buffer, BAPTA, in order to suppress the endogenous calcium-dependent chloride conductivity. Channels formed by α1G subunits demonstrated the following ratios of the amplitudes of macroscopic currents and potentials of zero current: I Ca:I Ba:I Sr = 1.00:0.75:1.12 and E CaE BaE Sr. For channels that were formed by α1H and α1I subunits, these ratios were as follows: I Ca:I Ba:I Sr = 1.00:1.20:1.17, E CaE BaE Sr and I Ca:I Ba:I Sr = 1.00:1.48: 1.45, E CaE BaE Sr respectively. The different macroscopic conductivities and similar potentials of zero current typical of α1G and α1I channels indicate that, probably, various bivalent cations can in a differential manner influence the stochastic parameters of functioning of these channels. At the same time, channels formed by α1H subunits are characterized by more positive potentials of zero current for Ca2+. It seems possible that the selectivity of the above channels is determined by mechanisms that mediate the selectivity of most high-threshold calcium channels (more affine binding of Ca2+ inside the pore). Neirofiziologiya/Neurophysiology, Vol. 37, No. 4, pp. 319–329, July–August, 2005.  相似文献   

5.
Phytate, the major source of phosphorus in seeds, exists as a complex with different metal ions. Alkaline phytases are known to dephosphorylate phytate complexed with calcium ions in contrast to acid phytases that act only on phytic acid. A recombinant alkaline phytase from Bacillus sp. MD2 has been purified and characterized with respect to the effect of divalent metal ions on the enzyme activity and stability. The presence of Ca2+ on both the enzyme and the substrate is required for optimal activity and stability. Replacing Ca2+ with Ba2+, Mn2+, Mg2+ and Sr2+ in the phytase resulted in the expression of > 90% of the maximal activity with calcium-phytate as the substrate, while Fe2+ and Zn2+ rendered the enzyme inactive. On the other hand, the calcium loaded phytase showed significant activity (60%) with sodium phytate and lower activity (17-20%) with phytate complexed with only Mg2+, Sn2+ and Sr2+, respectively. On replacing Ca2+ on both the enzyme and the substrate with other metal ions, about 20% of the maximal phytase activity was obtained only with Mg2+ and Sr2+, respectively. Only Ca2+ resulted in a marked increase in the melting temperature (Tm) of the enzyme by 12-21 °C, while Ba2+, Mn2+, Sr2+ or Cu2+ resulted in a modest (2-3.5 °C) increase in Tm. In the presence of 1-5 mM Ca2+, the optimum temperature of the phytase activity was increased from 40 °C to 70 °C, while optimum pH of the enzyme shifted by 0.4-1 pH unit towards the acidic region.  相似文献   

6.
Iron deficiency is the most prevalent micronutrient deficiency worldwide. Whereas dietary calcium is known to reduce the bioavailability of iron, the molecular basis of this interaction is not understood. We tested the hypothesis that divalent metal-ion transporter-1 (DMT1)—the principal or only mechanism by which nonheme iron is taken up at the intestinal brush border—is shared also by calcium. We expressed human DMT1 in RNA-injected Xenopus oocytes and examined its activity using radiotracer assays and the voltage clamp. DMT1 did not mediate 45Ca2+ uptake. Instead, we found that Ca2+ blocked the Fe2+-evoked currents and inhibited 55Fe2+ uptake in a noncompetitive manner (Ki ≈ 20 mM). The mechanism of inhibition was independent of voltage and did not involve intracellular Ca2+ signaling. The alkaline-earth metal ions Ba2+, Sr2+, and Mg2+ also inhibited DMT1-mediated iron-transport activity. We conclude that Ca2+ is a low-affinity noncompetitive inhibitor—but not a transported substrate—of DMT1, explaining in part the effect of high dietary calcium on iron bioavailability.  相似文献   

7.
This study was undertaken to elucidate the effect of the essential oil from Alpinia speciosa (EOAs) on cardiac contractility and the underlying mechanisms. The essential oil was obtained from Alpinia speciosa leaves and flowers and the oil was analyzed by GC-MS method. Chemical analysis revealed the presence of at least 18 components. Terpinen-4-ol and 1,8-cineole corresponded to 38% and 18% of the crude oil, respectively. The experiments were conducted on spontaneously-beating right atria and on electrically stimulated left atria isolated from adult rats. The effect of EOAs on the isometric contractions and cardiac frequency in vitro was examined. EOAs decreased rat left atrial force of contraction with an EC50 of 292.2 ± 75.7 μg/ml. Nifedipine, a well known L-type Ca2+ blocker, inhibited in a concentration-dependent manner left atrial force of contraction with an EC50 of 12.1 ± 3.5 μg/ml. Sinus rhythm was diminished by EOAs with an EC50 of 595.4 ± 56.2 μg/ml. Whole-cell L-type Ca2+ currents were recorded by using the patch-clamp technique. EOAs at 25 μg/ml decreased ICa,L by 32.6 ± 9.2% and at 250 μg/ml it decreased by 89.3 ± 7.4%. Thus, inhibition of L-type Ca2+ channels is involved in the cardiodepressive effect elicited by the essential oil of Alpinia speciosa in rat heart.  相似文献   

8.
We have previously elucidated a new paradigm for the metal ion-induced helix-helix assembly in the natural γ-carboxyglutamic acid (Gla)-containing class of conantokin (con) peptides, typified by con-G and a variant of con-T, con-T[K7Gla], independent of the hydrophobic effect. In these “metallo-zipper” structures, Gla residues spaced at i, i + 4, i + 7, i + 11 intervals, which is similar to the arrangement of a and d residues in typical heptads of coiled-coils, coordinate with Ca2+ and form specific antiparallel helical dimers. In order to evaluate the common role of Gla residues in peptide self-assembly, we extend herein the same Gla arrangement to designed peptides: NH2-(γLSγEAK)3-CONH2 (peptide 1) and NH2-γLSγEAKγLSγQANγLSγKAE-CONH2 (peptide 2). Peptide 1 and peptide 2 exhibit no helicity alone, but undergo structural transitions to helical conformations in the presence of a variety of divalent cations. Sedimentation equilibrium ultracentrifugation analyses showed that peptide 1 and peptide 2 form helical dimers in the presence of Ca2+, but not Mg2+. Folding and thiol-disulfide rearrangement assays with Cys-containing peptide variants indicated that the helical dimers are mixtures of antiparallel and parallel dimers, which is different from the strict antiparallel strand orientations of con-G and con-T[K7γGla] dimers. These findings suggest that the Gla arrangement, i, i + 4, i + 7, i + 11, i + 14, plays a key role in helix formation, without a strict adherence to strand orientation of the helical dimer.  相似文献   

9.
Using the standard voltage-clamp technique in the whole-cell mode, we studied the characteristics of barium currents (I Ba; Ba2+ concentration in the external solution was 5 mM) carried through L-type Ca2+ channels in the membrane of myocytes of the resistive mesenteric artery from normotensive and genetically hypertensive rats (NR and GHR, respectively). To perforate the membrane, we used amphotericin B. The arbitrary density of I Ba through the plasma membrane of GHR myocytes significantly exceeded this parameter in the NR group. For both animal groups, activation curves plotted as the dependence of the membrane conductance (G Ba) on the membrane potential were not significantly different: the membrane potential for half activation (V 0.5) of I Ba in the NR myocytes was equal to 1.0 ± 0.3 mV with slope factor k = 6.3 ± 0.4 mV, whereas in the GHR myocytes V 0.5 = -1.6 ± 0.2 mV and k = 6.2 ± 0.5 mV. The stationary inactivation curves for I Ba differed significantly: in the NR myocytes, V 0.5 = -24.2 ± 0.4 mV and k = 8.3 ± 0.2 mV, whereas in the GHR myocytes such parameters were, respectively, -21.4 ± 0.4 and 8.7 ± 0.3 mV. The pattern of intersection of stationary activation and stationary inactivation curves for I Ba was indicative of the existence of a window current, i.e., the non-inactivating component of I Ba within the -40 to ±20 mV range; the phenomenon was clearly pronounced in the GHR myocytes. Differences in the arbitrary density of integral I Ba and window current were observed. These differences can cause an increased tone of the blood vessels in hypertensive animals.  相似文献   

10.
The gene for a novel cation/H+ antiporter from Puccinellia tenuiflora, PutCAX1, was cloned from a cDNA library. The PutCAX protein was localized in the vacuolar membrane using a GFP marker. Several yeast transformants were created using full-length and truncated form of PutCAX1 and their growths in the presence of various cations (Mg2+, Ca2+, Mn2+, Ni2+, Cu2+, Zn2+, Se2+, and Ba2+) were analyzed. PutCAX1 expression was found to affect the response to Ca2+ and Ba2+ in yeast. The PutCAX1 and C-terminally truncated PutCAX1 (ΔCPutCAX1) transformants grew in the presence of 70 mM Ca2+ as well as in the presence of 8 mM Ba2+. However, the ΔCPutCAX1 transformant was able to grow in the presence of 20 mM Ba2+ while the PutCAX1 transformant could not. On the other hand, expression of the N-terminally truncated form and the N- and C-terminally truncated form failed to suppress the Ca2+ or Ba2+ sensitivity of yeast. These results suggest that PutCAX1 can complement the active Ca2+ transporters at some level and confer yeast Ba2+ tolerance, and that the N- and C-terminal regions of PutCAX1 play important roles in increasing the Ca2+ or Ba2+ tolerance of yeast.  相似文献   

11.
12.
Complexation of d-gluconate (Gluc) with Ca2+ has been investigated via 1H, 13C and 43Ca NMR spectroscopy in aqueous solutions in the presence of high concentration background electrolytes (1 M ? I ? 4 M (NaCl) ionic strength). From the ionic strength dependence of its formation constant, the stability constant at 6 ? pH ? 11 and at I → 0 M has been derived (). The protonation constant of Gluc at I = 1 M (NaCl) ionic strength was also determined and was found to be log Ka = 3.24 ± 0.01 (13C NMR) and log Ka = 3.23 ± 0.01 (1H NMR). It was found that 1H and 13C NMR chemical shifts upon complexation (both with H+ and with Ca2+) do not vary in an unchanging way with the distance from the Ca2+/H+ binding site. From 2D 1H-43Ca NMR spectra, simultaneous binding of Ca2+ to the alcoholic OH on C2 and C3 was deduced. Molecular modelling results modulated this picture by revealing structures in which the Gluc behaves as a multidentate ligand. The five-membered chelated initial structure was found to be thermodynamically more stable than that derived from a six-membered chelated initial structure.  相似文献   

13.
As part of an effort to inhibit S100B, structures of pentamidine (Pnt) bound to Ca2+-loaded and Zn2+,Ca2+-loaded S100B were determined by X-ray crystallography at 2.15 Å (Rfree = 0.266) and 1.85 Å (Rfree = 0.243) resolution, respectively. These data were compared to X-ray structures solved in the absence of Pnt, including Ca2+-loaded S100B and Zn2+,Ca2+-loaded S100B determined here (1.88 Å; Rfree = 0.267). In the presence and absence of Zn2+, electron density corresponding to two Pnt molecules per S100B subunit was mapped for both drug-bound structures. One Pnt binding site (site 1) was adjacent to a p53 peptide binding site on S100B (± Zn2+), and the second Pnt molecule was mapped to the dimer interface (site 2; ± Zn2+) and in a pocket near residues that define the Zn2+ binding site on S100B. In addition, a conformational change in S100B was observed upon the addition of Zn2+ to Ca2+-S100B, which changed the conformation and orientation of Pnt bound to sites 1 and 2 of Pnt-Zn2+,Ca2+-S100B when compared to Pnt-Ca2+-S100B. That Pnt can adapt to this Zn2+-dependent conformational change was unexpected and provides a new mode for S100B inhibition by this drug. These data will be useful for developing novel inhibitors of both Ca2+- and Ca2+,Zn2+-bound S100B.  相似文献   

14.
The Iberian Peninsula encompasses more than 80% of the species richness of European aquatic ranunculi. The floristic diversity of the phytocoenosis characterised by aquatic Ranunculus and the main physical–chemical factors of the water were studied in 43 localities of the central Iberian Peninsula. Four aquatic Ranunculus communities are found in most of the aquatic environments. These are species-poor and have an uneven distribution: three species of Batrachium are heterophyllous and their communities are distributed in different aquatic ecosystems on silicated substrates; one species is homophyllous and its community occurs in various aquatic ecosystems with carbonated waters. In the Mediterranean climate, Ranunculus species are present in different habitats, as shown by the results of all the statistical analyses. Ranunculus trichophyllus communities occur in base-rich waters with a high buffering capacity (2273.44 ± 794.57 mg CaCO3 L−1) and a high concentration of cations (Ca2+, 121 ± 33.12 mg L−1; Mg2+, 71.64 ± 82.77 mg L−1), nitrates (2.89 ± 4.80 mg L−1), ammonium (2.19 ± 1.36 mg L−1) and sulphates (216.25 ± 218.54 mg L−1). Ranunculus penicillatus communities are found in flowing waters with a high concentration of phosphates (0.48 ± 0.6 mg L−1) and intermediate buffering capacity (683.66 ± 446.76 mg CaCO3 L−1). Both Ranunculus pseudofluitans and Ranunculus peltatus communities grow in waters with low buffering capacity (R. pseudofluitans, 385.91 ± 209.2 mg CaCO3 L−1; R. peltatus, 263.3 ± 180.36 mg CaCO3 L−1), and a low concentration of cations (R. pseudofluitans: Ca2+, 12.57 ± 9.42 mg L−1; Mg2+, 3.42 ± 1.67 mg L−1; R. peltatus: Ca2+, 15 ± 18.26 mg L−1; Mg2+, 6.26 ± 8.89 mg L−1) and nutrients (R. pseudofluitans: nitrates, 0.23 ± 0.2 mg L−1; phosphates, 0.09 ± 0.1 mg L−1; R. peltatus: nitrates, 0.19 ± 0.21 mg L−1; phosphates, 0.09 ± 0.12 mg L−1); the first in flowing waters, the latter in still waters.  相似文献   

15.
Slow inactivated states in voltage-gated ion channels can be modulated by binding molecules both to the outside and to the inside of the pore. Thus, external K+ inhibits C-type inactivation in Shaker K+ channels by a “foot-in-the-door” mechanism. Here, we explore the modulation of a very long-lived inactivated state, ultraslow inactivation (IUS), by ligand binding to the outer vestibule in voltage-gated Na+ channels. Blocking the outer vestibule by a mutant μ-conotoxin GIIIA substantially accelerated recovery from IUS. A similar effect was observed if Cd2+ was bound to a cysteine engineered to the selectivity filter (K1237C). In K1237C channels, exposed to 30 μM Cd2+, the time constant of recovery from IUS was decreased from 145.0 ± 10.2 s to 32.5 ± 3.3 s (P < 0.001). Recovery from IUS was only accelerated if Cd2+ was added to the bath solution during recovery (V = −120 mV) from IUS, but not when the channels were selectively exposed to Cd2+ during the development of IUS (−20 mV). These data could be explained by a kinetic model in which Cd2+ binds with high affinity to a slow inactivated state (IS), which is transiently occupied during recovery from IUS. A total of 50 μM Cd2+ produced an ∼8 mV hyperpolarizing shift of the steady-state inactivation curve of IS, supporting this kinetic model. Binding of lidocaine to the internal vestibule significantly reduced the number of channels entering IUS, suggesting that IUS is associated with a conformational change of the internal vestibule of the channel. We propose a molecular model in which slow inactivation (IS) occurs by a closure of the outer vestibule, whereas IUS arises from a constriction of the internal vestibule produced by a widening of the selectivity filter region. Binding of Cd2+ to C1237 promotes the closure of the selectivity filter region, thereby hastening recovery from IUS. Thus, Cd2+ ions may act like a foot-on-the-door, kicking the IS gate to close.  相似文献   

16.
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   

17.
In present study, an HPLC method coupled with photodiode array detector (HPLC-PDA) was established for determination and pharmacokinetics of gastrodin (GAS) in human plasma after an oral administration of GAS capsule. In the method, ethanol and dichloromethane were respectively used for deproteinization and purification during the sample preparation procedure. Separation of GAS was achieved on an AichromBond-AQ C18 column (5 μm, 150 mm × 4.6 mm) with the mobile phase of methanol–0.1% phosphoric acid solution (2:98, v/v) at a flow rate of 0.8 ml/min. The wavelength was set at 220 nm and the injection volume was 20 μl. Under the conditions, the calibration curve was linear within the concentration range of 50–4000 ng/ml with the correlation coefficient (r) of 0.99554 (weight = 1/X2) and the lower limit of quantification (LLOQ) was 50 ng/ml. The inter- and intra-day precisions were less than 11% and the accuracies (%) were within the range of 95.55–103.78%. The extraction recoveries were over 65% with RSDs less than 5.50%. The GAS was proved to be stable under tested conditions. Thus, the method was valid enough to be applied for pharmacokinetic study of GAS in human plasma. The pharmacokinetic parameters of GAS in human plasma after an oral administration of 200 mg GAS capsule were described as: Cmax, 1484.55 ± 285.05 ng/ml; Tmax, 0.81 ± 0.16 h; t1/2α, 3.78 ± 2.33 h; t1/2β, 6.06 ± 3.20 h; t1/2Ka, 0.18 ± 0.53 h; K12, 0.18 ± 0.41/h; K21, 0.20 ± 0.16/h; K10, 4.11 ± 15.81/h; V1/F, 180.35 ± 89.44 L; CL/F, 62.50 ± 140.03 l/h; AUC0→t, 5619.41 ± 1972.88 (ng/ml) h; and AUC0→∞, 7210.26 ± 3472.74 (ng/ml) h, respectively. These will be useful for the clinical application of GAS.  相似文献   

18.
The present study reports a detailed investigation into the interaction of [Cr(phen)2(dppz)]3+ and [Cr(phen)3]3+ with transferrin, the key protein for the transport of Fe3+ in blood plasma; its cycle holds promise as an attractive system for strategies of drug targeting to tumor tissues. This can allow us to understand further the role of both complexes as sensitizers in photodynamic therapy (PDT). Chromium(III) complexes, [Cr(phen)2(dppz)]3+ and [Cr(phen)3]3+, (phen = 1,10-phenanthroline and dppz = dipyridophenazine), where dppz is a planar bidentate ligand with an extended π system, have been found to bind strongly with apotransferrin (apoTf) with an intrinsic binding constant, Kb, of (1.8 ± 0.3) × 105 M− 1 and (1.1 ± 0.1) × 105 M− 1 at 299 K, for apoTf-[Cr(phen)2(dppz)]3+ and apoTf-[Cr(phen)3]3+, respectively. The interactions of apoTf with the different Cr(III) complexes were assessed employing UV-visible absorption, fluorescence and circular dichroism spectroscopy. The relative fluorescence intensity of the protein decreased when the increasing concentration of Cr(III) complex was added, suggesting that perturbation around the Trp and Tyr residues took place. The analysis of the thermodynamic parameters ΔG, ΔH, ΔS indicated that the presence of the Cr(III) complex stabilizes the protein with a strong entropic contribution. The binding distances and transfer efficiencies for apoTf-[Cr(phen)2(dppz)]3+ and apoTf-[Cr(phen)3]3+ binding reactions were calculated according to Föster theory of non-radiation energy transfer. All these experimental results suggest that [Cr(phen)2(dppz)]3+ and [Cr(phen)3]3+ bind strongly to apoTf indicating that this protein could act as a carrier of these complexes for further applications in PDT.  相似文献   

19.
We have obtained evidence that the Ca2+-selective current activated by Ca2+ store depletion (Ca2+ release-activated Ca2+ current; I crac) in Jurkat T lymphocytes is augmented in a time-dependent manner by Ca2+ itself. Whole cell patch clamp experiments employed high cytosolic Ca2+-buffering conditions to passively deplete Ca2+ stores. Rapidly switching to nominally Ca2+-free extracellular buffer instantaneously reduced I crac measured at −100 mV to leak current level. Unexpectedly, readmission of 2 mm Ca2+ instantaneously restored only 38 ± 5% (mean ±sem; n = 9) of the full I crac amplitude. The remainder reappeared in a monotonic time-dependent manner over 10 to 20 sec. Rapid vs. slow intracellular Ca2+ chelators did not alter this process, and inorganic I crac blockers did not regenerate it, arguing against an intracellular site of action. The effect was specific to Ca2+: introduction of the permeant ions, Ba2+ or Sr2+, failed to invoke time-dependent I crac reappearance. Moreover, equimolar substitution of Ba2+ for Ca2+ initially produced Ba2+ current of similar magnitude to the full Ca2+ current, but the Ba2+ current decayed monotonically to <50% of its initial amplitude in <20 sec. Conversely, return to Ca2+ produced a time-dependent increase in I crac to its larger Ca2+ permeation level. Thus Ca2+ appears to selectively promote a reversible transition of I crac that results in larger current flux, and at least partially explains the selectivity of this current for Ca2+ over other divalent ions. Received: 30 August 1995/Revised: 7 November 1995  相似文献   

20.
Octopamine plays an important role in mediating reward signals in olfactory learning and memory formation in insect. However, its target molecules and signaling pathways are still unknown. In this study, we investigated the effects of octopamine on the voltage-activated Ca2+ channels expressed in native Kenyon cells isolated from the mushroom body of the cricket (Gryllus bimaculatus) brain. The cell-attached patch clamp recordings with 100 mM Ba2+ outside showed the presence of dihydropyridine (DHP) sensitive L-type Ca2+ channels with a single channel conductance of approximately 21 ± 2 pS (n = 12). The open probability (NPo) of single Ca2+ channel currents decreased by about 29 ± 7% (n = 6) by bath application of 10 μM octopamine. Octopamine-induced decrease in Po was imitated by bath application of 8-Br-cAMP, a membrane-permeable cAMP analog. Pre-treatment of Kenyon cells with the octopamine receptor antagonist phentolamine blocked the inhibitory effect of octopamine on Ca2+ channels. Pre-treatment of Kenyon cells with H-89, a selective inhibitor of cAMP-dependent protein kinase (PKA) attenuated the inhibitory effect of bath applied octopamine on Ca2+ channels. These results indicate that DHP-sensitive L-type Ca2+ channel is a target protein for octopamine and its modulation is mediated via cAMP and PKA-dependent signaling pathways in freshly isolated Kenyon cell in the cricket G. bimaculatus.  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