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.
Thio-heterocyclic fused naphthalimides with aminoalkyl side chains were designed, synthesized and evaluated. These compounds have long wavelength absorptions and binding affinities to Calf thymus DNA. They could photodamage supercoiled pBR322 DNA from form I (closed) to II (nicked) at a concentration as low as 0.5 microM and to form III (linear) at a concentration of 50 microM. A possible mechanism of superoxide anion was provided.  相似文献   

2.
The neutrophil enzyme myeloperoxidase catalyzes the oxidation of tyrosine to tyrosyl radicals, which cross-link to proteins and initiate lipid peroxidation. Tryptophan is present in plasma at about the same concentration as tyrosine and has a similar one-electron reduction potential. In this investigation, we have determined the ability of myeloperoxidase to catalyze the oxidation of tryptophan to assess whether or not this reaction may contribute to oxidative stress at sites of inflammation. We show that tryptophan is a poor substrate for myeloperoxidase because, even though it reacts rapidly with compound I (kI 2.1 x 10(6) M(-1)s(-1)), it reacts sluggishly with compound II (kII 7 M(-1)s(-1)). Tryptophan reversibly inhibited production of hypochlorous acid by purified myeloperoxidase by converting the enzyme to a mixture of compound II and compound III. It gave 50% inhibition (I50) at a concentration of 2 microM. In contrast, it was an ineffective inhibitor of hypochlorous acid production by human neutrophils (I50 80 microM) unless superoxide dismutase was present (I50 5 microM). We propose that compound I of myeloperoxidase will oxidize tryptophan at sites of inflammation. Enzyme turnover will result from the reaction of superoxide or tyrosine with compound II. Thus, tryptophan radicals are potential candidates for exacerbating oxidative stress during inflammation.  相似文献   

3.
Prion diseases are fatal neurodegenerative disorders that are caused by the conversion of a normal host-encoded protein, PrP(C), to an abnormal, disease-causing form, PrP(Sc). This paper reports that cyclodextrins have the ability to reduce the pathogenic isoform of the prion protein PrP(Sc) to undetectable levels in scrapie-infected neuroblastoma cells. Beta-cyclodextrin removed PrP(Sc) from the cells at a concentration of 500 microM following 2 weeks of treatment. Structure activity studies revealed that antiprion activity was dependent on the size of the cyclodextrin. The half-maximal inhibitory concentration (IC(50)) for beta-cyclodextrin was 75 microM, whereas alpha-cyclodextrin, which possessed less antiprion activity, had an IC(50) of 750 microM. This report presents cyclodextrins as a new class of antiprion compound. For decades, the pharmaceutical industry has successfully used cyclodextrins for their complex-forming ability; this ability is due to the structural orientation of the glucopyranose units, which generate a hydrophobic cavity that can facilitate the encapsulation of hydrophobic moieties. Consequently, cyclodextrins could be ideal candidates for the treatment of prion diseases.  相似文献   

4.
The effect of protein kinase C (PKC) inhibitors on porcine oocyte activation by calcium ionophore A23187 was studied. Calcium ionophore applied in a 50 microM concentration for 10 min induced activation in 74% of oocytes matured in vitro. When the ionophore-treated oocytes were exposed to the effect of bisindolylmaleimide I, which inhibits calcium-dependent PKC isotypes (PKC-alpha, -beta(I), -beta(II), -gamma,) and calcium-independent PKC isotypes (PKC-delta, -epsilon), the portion of activated oocytes decreased (at a concentration of 100 nM, 2% of the oocytes were activated). Go6976, the inhibitor of calcium-dependent PKC isotypes PKC-alpha, -beta(I) did not prevent the action of the oocytes treated with calcium ionophore in concentrations from 1 to 100 microM. The inhibitor of PKC-beta(I) and beta(II) isotypes, hispidin, in a concentration of 2 microM-2 mM, was not effective either. The inhibitor of PKC-delta isotype, rottlerin, suppressed activation of the oocytes by calcium ionophore (no oocyte was activated at 10 microM concentration). The PKC-delta isotype in matured porcine oocytes, studied by Western blot analysis, appeared as non-truncated PKC-delta of 77.5 kDa molecular weight, on the one hand, and as truncated PKC-delta, which was present in the form of a doublet of approximately 62.5 and 68 kDa molecular weight, on the other hand. On the basis of these results, it can be supposed that PKC participates in the regulation of processes associated with oocyte activation. Calcium-dependent PKC-alpha, -beta isotypes do not seem to play any significant role in calcium activation. The activation seems to depend on the activity of the calcium-independent PKC-delta isoform.  相似文献   

5.
Dopamine D2 receptors (D2Rs; short form, which is one of the alternative splicing variants) expressed in COS-7 cells are internalized in an agonist-dependent manner only when G protein-coupled receptor kinase 2 (GRK2) is coexpressed [Ito, K., Haga, T., Lameh, J. & Sadée, W., (1999) Eur. J. Biochem. 260, 112-119]. We have examined the effects of coexpression of dynamin, a small molecular mass GTP-binding protein, rab5A, and their mutants on the internalization of D2Rs in the presence of both dopamine (10 or 100 microM) and GRK2. The rate and extent of D2R internalization was increased or decreased by coexpression of dynamin I or a dominant-negative form of dynamin I (dynamin I K44E), respectively. The effects of coexpressing these two dynamins were more prominent at 10 microM dopamine than at 100 microM. In the presence of 10 microM dopamine, internalization of D2R was completely suppressed when dynamin I K44E was coexpressed, and the half-life (t 1/2) of D2R internalization decreased relative to cells not expressing dynamin from 82 to 29 min when dynamin I was coexpressed. Internalization of D2Rs was facilitated or suppressed by coexpression of a constitutively active form of rab5A (rab5A Q79L) or a dominant-negative form of rab5A (rab5A S34N), respectively. The t 1/2 of D2R internalization at 10 microM dopamine decreased from 82 to 16 min in cells coexpressing rab5A Q79L. The effect of coexpression of rab5A S34N was more apparent at 100 microM dopamine than at 10 microM; the t 1/2 of D2R internalization at 100 microM dopamine increased from 20 to 56 min and the proportion of internalized D2Rs after 120 min decreased from 53 to 28%. These results indicate that the internalization of D2Rs is dependent on the action of dynamin as well as GRK2, and is regulated by the action of rab5A.  相似文献   

6.
The fluorescent nucleotide 2',3'-trinitrophenyl-ATP (TNP-ATP) binds at the triphosphate substrate binding site of the large (Klenow) fragment of DNA polymerase I (Pol I) as detected by direct binding studies measuring the increase in fluorescence of this ligand (n = 1.0, KD = 0.07 microM). The enzyme-TNP-ATP complex binds Mg2+ and Mn2+ tightly (KD = 0.05 microM) as measured by an increase in fluorescence on titrating with these metals. The substrate dGTP competitively displaces TNP-ATP from the enzyme (KD = 5.7 microM) de-enhancing the fluorescence. The polymerase reaction is half-maximally inhibited by 0.8 microM TNP-ATP in the presence of dATP (10 microM) as substrate. A region of the amino acid sequence of Pol I (peptide I) consisting of residues 728-777 has been synthesized and found to contain significant secondary structure by CD both in water and 50% methanol/water. In water at 3 degrees C, peptide I binds the substrate analog TNP-ATP (KD = 0.03 microM) with a stoichiometry of 0.2. In 50% methanol at 3 degrees C, peptide I binds TNP-ATP with a higher stoichiometry than in water, consistent with a 1:1 complex, but biphasically (16% of the peptide, KD = 0.09 microM; 84% of the peptide, KD = 5.0 microM), and competitively binds the Pol I substrates dATP, TTP, and dGTP (KD = 230-570 microM). Evidence from size exclusion high performance liquid chromatography suggests that these two forms of the peptide are monomer and dimer, respectively. Significantly, the peptide I-TNP-ATP complex binds duplex DNA, tightly (KD = 0.1-0.5 microM) and stoichiometrically, and single stranded DNA more weakly. The peptide I-duplex DNA complex binds both TNP-ATP (KD = 0.5-1.5 microM) and Pol I substrates (KD = 350-2100 microM) stoichiometrically. In a control experiment, a second peptide, peptide II, based on residues 840-888 of the Pol I sequence, retains secondary structure, as detected by CD, but displays no binding of TNP-ATP. The ability of peptide I, which represents only 8% of the large fragment of Pol I, to bind both substrates and duplex DNA indicates that residues 728-777 constitute a major portion of the substrate binding site of this enzyme.  相似文献   

7.
The novel DNA interactive isoquinolino[5,4-ab]phenazine derivatives were designed and synthesized. Their inhibitory abilities toward topoisomerase I, antitumor activities and DNA photo-cleaving abilities were examined. The substituents at peri sites of two phenazine N atoms played very important roles for all these biological activities. At a concentration of 100 microM, all these phenazine derivatives (but A2 and A6) exhibited an inhibitory activity toward topoisomerase I. A6 had efficient antitumor activities against both human lung cancer cell (A549) and murine leukemia cell (P388). A1, A5, and A6 exhibited antitumor activities selectively against P388. A2 was the most efficient DNA photocleaver, which had converted supercoiled DNA from form I to form II at <1 microM. Under anaerobic conditions, the electron transfer mechanism mainly contributed to DNA photo-induced cleavage, while under aerobic conditions, superoxide anion was also involved in this process.  相似文献   

8.
In a cytosolic extract from rat liver, the number and the concentration of ADP-binding sites as well as their dissociation constants were determined by using the rate-of-dialysis technique. Interfering cytosolic adenylate kinase was extracted from the cytosol by affinity chromatography on Ap5A-agarose, and remaining traces of enzyme activity were inhibited with (+)-catechin. Binding of ADP to cytosolic proteins was increased by poly(ethylene glycol) and decreased by EDTA. The effect of 0.1 mM-EDTA could be reversed by addition of equimolar concentrations of Mn2+ or Mg2+. In presence of 5% poly(ethylene glycol), added to increase local protein concentration, two binding sites for ADP were observed, with KD values of 1.9 microM (site I) and 10.8 microM (site II). The concentration of these binding sites, when extrapolated to cellular protein concentrations, were 30 microM (site I) and 114 microM (site II). It is concluded that a minimum of about 50% of total cytosolic ADP is bound to proteins, and that the ratio of free ATP/free ADP is at least twice that of total ATP/total ADP.  相似文献   

9.
A recent paper (Buchberger, W., 1988, J. Chromatogr. 432, 57) on lactoperoxidase-catalyzed bromination of tyrosine and thyroglobulin stated, without evidence, that thyroid peroxidase (TPO) is able to use bromide as a substrate. This was in disagreement with unpublished experiments previously performed in this laboratory, and we undertook, therefore, to examine this subject further. Highly purified porcine TPO was compared with lactoperoxidase (LPO) and chloroperoxidase (CPO) for ability to catalyze bromination of tyrosine, thyroglobulin, and bovine serum albumin (BSA). The incubation mixture contained 50-100 nM peroxidase, 10-500 microM 82Br-, tyrosine (150 microM), thyroglobulin (0.3 or 1 microM), or BSA (7.5 microM), and a source of H2O2. The latter was either generated by glucose (1 mg/ml)-glucose oxidase (0.5 or 1 micrograms/ml), or added initially as a bolus (100 microM). With TPO, formation of organically bound 82Br was undetectable under all conditions in the pH range 5.4-7.0. Lactoperoxidase and CPO, on the other hand, displayed considerable brominating activity. Lactoperoxidase was much more active at pH 5.4 than at pH 7.0 and was more active with BSA as acceptor than with tyrosine or thyroglobulin. The distribution of 82Br among the various amino acids in LPO-brominated thyroglobulin and BSA was determined by HPLC. As expected, monobromotyrosine and dibromotyrosine together comprised the greatest part of the bound 82Br. However, a surprisingly high percentage (20-25%) was present as monobromohistidine. Evidence was also obtained for the presence of a small percentage of the bound 82Br as tetrabromothyronine. Peroxidase-catalyzed bromination probably depends on the oxidation of Br- to Br+ by the Compound I form of the enzyme. Since oxidation of Br- to Br+ requires a stronger oxidant than oxidation of I- to I+, our results suggest that Compound I of LPO and of CPO has a higher oxidation potential than Compound I of TPO. In vivo experiments with rats on a low iodine diet injected with 82Br- showed that even under conditions of high stimulation by thyrotropic hormone, there is negligible formation of organic bromine in the thyroid. Measurements of thyroid:serum concentration ratios for 82Br- in similar rats provided no evidence that Br- is a substrate for the iodide transport system of the thyroid.  相似文献   

10.
B Singer  F Chavez  S J Spengler 《Biochemistry》1986,25(6):1201-1205
O4-Methyl-, O4-ethyl-, and O4-isopropylthymidine 5'-triphosphates, which can be formed by N-nitroso carcinogens, were tested for their ability to substitute for thymidine 5'-triphosphate (dTTP) in synthesis catalyzed by Escherichia coli DNA polymerase I (Pol I) by using activated DNA or synthetic polymers as templates. All could substitute for dTTP for short periods, the rate and extent decreasing with the size of the alkyl group. Because the structure of O4-alkylthymidine does not permit normal hydrogen bond formation with deoxyadenosine, it was inferred that eventual formation of a poor or frayed primer end was responsible for termination of synthesis. Synthesis of polymers at temperatures ranging from 0 to 40 degrees C showed that the extent of incorporation using the O4-alkyl-dTTPs was favored, relative to dTTP, when the terminal helical structure was stabilized by low temperatures. Kmapp values were determined for each O4-alkyldeoxynucleoside 5'-triphosphate. These values were 0.7 microM for dTTP, 5 microM for methyl-dTTP, 11 microM for ethyl-dTTP, and 33 microM for isopropyl-dTTP. O4-Alkyl-dTTPs were tested for their ability to inhibit or compete with dTTP incorporation and found to have a minimal effect, even when present at high concentration. These experiments indicated that Pol I can incorporate deoxynucleotides with O4-alkyl substituents into an ordered DNA structure. A postulated base-pairing scheme with deoxyadenosine is described.  相似文献   

11.
G Soslau  J Parker    J W Nelson 《Nucleic acids research》1986,14(18):7237-7252
These studies employed the synthetic linear DNA, poly dGdC, in the B and cobalt hexammine chloride (Co)-induced Z form to determine the effect of conformation on protein-DNA interactions. The rate of the reaction of the restriction endonucleases, Hha I and Cfo I, are reduced with Z DNA as compared to B DNA. The ability of both restriction endonucleases to react with an aggregate form of Z DNA (Z* DNA) is found to depend upon how the Z* DNA is formed. When Z* DNA is induced by low concentrations of Co (50 microM), the endonucleases remain active. In the presence of 100 microM Co, which causes increased aggregation, the endonucleases are inactive. The Hha I DNA methyltransferase reacts at equal rates with the B, Z and low cobalt Z* forms and at a greatly reduced rate with the high cobalt Z* form. These results are significantly different than those observed with Z form dGdC tracts inserted into circular DNA molecules.  相似文献   

12.
3 beta-Hydroxy-5 alpha-cholest-8(14)-en-15-one (I) and (25R)-26-hydroxycholesterol (II), both potent regulators of sterol biosynthesis, have been found to show synergism in the reduction of the levels of HMG-CoA reductase activity in CHO-K1 cells. When equimolar concentrations of I and II were added in combination, synergistic reduction (p less than 0.0001) of enzyme activity was observed at total oxysterol concentrations of 0.1 microM, 0.2 microM, and 0.5 microM. Maximal synergistic effect in the lowering of reductase activity (28% greater than predicted) was observed at 0.1 microM total oxysterol concentration. Five additional experiments conducted with 50 nM oxysterols confirmed the synergistic effect at 0.1 microM total sterol concentration. These results suggest that the in vivo importance of I and II may be greater than that anticipated on the basis of the concentrations of the individual sterols.  相似文献   

13.
Polycationic molecules were studied either for their ability to displace the binding of basic fibroblast growth factor (bFGF) to high- and low-affinity membrane interaction sites and/or to modulate bFGF-induced proliferation of fibroblasts. Heparin-binding polypeptides, such as polylysine, protamine, histones, and thrombin-displaced [125I]bFGF bound to bovine brain membrane receptors. The most displacing polypeptides were those with the strongest affinity to heparin. Two of these polypeptides, protamine and polylysine, inhibited (at 5 microM) by more than 90% the mitogenic effect induced by bFGF on Chinese hamster lung fibroblast cells (CCL39). At the same dose, no effect was observed with basic proteins that do not bind to heparin, such as cytochrome C and lysozyme. An interesting observation was that protamine at 1 microM potentiated by 1.5-fold the mitogenic activity of bFGF, while it acted as an inhibitor at higher concentration.  相似文献   

14.
The enzyme, tyrosinase, was immobilized inside carbon paste electrodes (CPE) for the analysis of thiol-containing compounds such as the reduced form of glutathione (GSH) and L-cysteine. The measuring principle of this sensor is based on the blocking of the substrate recycling process between the enzyme and the electrode. The current response is monitored at -0.050 V versus Ag/AgCl. At this low potential, interferences from easily oxidizable species such as ascorbic acid and uric acid are minimized. The tyrosinase CPE is characterized both in steady state experiments and by flow injection analysis (FIA). GSH is used as the model thiol-containing compound for the study. The highest response for GSH was obtained around pH 6.5. A detection limit of 100 nM and 1 microM is achieved for GSH in steady state and in flow measurements, respectively. The analytical range for GSH is dependent on the concentration of the tyrosinase substrate (catechol). In steady state experiments, and at a lower substrate concentration (10 microM catechol), a linear range of 1-8 microM is found for GSH as compared with 5-30 microM at a higher substrate concentration of 20 microM catechol. Current response of the tyrosinase CPE is not affected by the oxidized form of GSH and L-cysteine (glutathione disulfide, GSSG, and L-cystine, respectively) and sulfur-containing compound such as methionine. The tyrosinase CPE can also detect coenzyme A, which makes it possible to construct biosensors based on enzymes producing or utilizing coenzyme A.  相似文献   

15.
Fructose-1,6-bisphosphatase (FBPase; EC 3.1.3.11) is strongly inhibited by AMP in vitro and, therefore, at physiological concentrations of substrate and AMP, FBPase should be completely inhibited. Desensitization of rabbit muscle FBPase against AMP inhibition was previously observed in the presence of rabbit muscle aldolase. In this study, we analysed the kinetics of an FBPase catalyzed reaction and interaction between chicken muscle FBPase and chicken muscle aldolase. The initial rate of FBPase reaction vs. substrate concentration shows a maximum activity at a concentration of 20 microM Fru-1,6P2 and then decreases. Assuming rapid equilibrium kinetics, the enzyme-catalyzed reaction was described by the substrate inhibition model, with Ks approximately 5 microM and Ksi approximately 39 microM and factor beta approximately 0.2, describing change in the rate constant (k) of product formation from the ES and ESSi complexes. Based on ultracentrifugation studies, aldolase and FBPase form a hetero-complex with approximately 1:1 stoichiometry with a dissociation constant (Kd) of 3.8 microM. The FBPase-aldolase interaction was confirmed via fluorescence investigation. The aldolase-FBPase interaction results in aldolase fluorescence quenching and its maximum emission spectrum shifting from 344 to 356 nm. The Kd of the FBPase-aldolase complex, determined on the basis of fluorescence changes, is 0.4 microM at 25 degrees C with almost 1:1 stoichiometry. This interaction increases the I(0.5) for the AMP inhibition of FBPase threefold, and slightly affects FBPase affinity to magnesium ions, increasing the Ka and Hill coefficient (n). No effect of aldolase on the FBPase pH optimum was observed. Thus, the decrease in FBPase sensitivity to AMP inhibition enables FBPase to function in vivo thanks to aldolase.  相似文献   

16.
We used a gel filtration method and a stable prostaglandin I2 (prostacyclin) analogue, iloprost, to study the kinetics of prostaglandin I2 binding by human serum proteins. Binding equilibrium experiments conducted at physiological prostaglandin I2 concentration (nM) yielded a KD of 10(-9) and a capacity of approx. 50 nM for the serum binding protein(s). Kinetic measurements gave a dissociation rate constant of 10(-3) s-1. When binding equilibrium was established at various ligand concentrations ranging from nM to microM, a result indicating an unsaturable binding was obtained utilizing this method. On the other hand, saturation was achieved with a ligand concentration as high as 50-100 microM by another binding method. A KD of 7 X 10(-5) and a capacity of approx. 600 microM was obtained. This apparent discrepancy was resolved by performing parallel experiments using purified human serum albumin samples and serum. It is concluded that the large quantity of serum albumin, approx. 600 microM, in serum may compensate for its low KD (approx. 10(-5] for prostaglandin I2, thus simulating a binding protein with a KD of 10(-9) and a limited capacity. These data offer direct information regarding how prostaglandin I2 is stabilized by serum and is transported to the platelet prostaglandin I2 receptors. There is a strong implication that serum albumin is the major if not the only protein responsible for binding of prostaglandin I2.  相似文献   

17.
A new heterocyclic family of (2-(dimethylamino)ethyl)-2-substituted phenylnaphtho[2,1-d]thiazole-5-carboxamides modified from naphthalimides was designed, synthesized, and quantitatively evaluated as antitumor agents and photonucleases. All these compounds were found to be more cytotoxic against P388 than against A549. B(3) (m-NO(2)) was found to be the strongest inhibitor for P388 with IC(50) of 1.49 microM, while B(2) was the most cytotoxic compound against A549 with IC(50) of 12 microM. B(4) (p-CH(3)), the most efficient DNA photocleaver, showed detectable DNA cleavage at 0.5 microM and total cleavage from form I to 100% form II at 50 microM. The photocleaving mechanism was changed with the modification to be via superoxide anion and radical.  相似文献   

18.
Under physiological conditions, small amounts of free arachidonic acid (AA) are released from membrane phospholipids, and cyclooxygenase (COX) and acyl-CoA synthetase (ACS) competitively act on this fatty acid to form prostaglandins (PGs) and arachidonoyl-CoA (AA-CoA). In the present study, we investigated the effects of eicosapentaenoic acid (EPA) and 15-hydroperoxyeicosapentaenoic acid (15-HPEPE) on the PG and AA-CoA formations from high and low concentrations of AA (60 and 5 microM) in rabbit kidney medulla microsomes. The kidney medulla microsomes were incubated with 60 or 5 microM [(14)C]-AA in 0.1M Tris/HCl buffer (pH 8.0) containing cofactors of COX (reduced glutathione and hydroquinone) and cofactors of ACS (ATP, MgCl(2) and CoA). After incubation, PG (as total PGs), AA-CoA and residual AA were separated by selective extraction using petroleum ether and ethyl acetate. EPA reduced the PG and AA-CoA formations from both 60 and 5 microM AA. In contrast, 15-HPEPE decreased the PG formation without affecting the AA-CoA formation from 60 microM AA, and increased the AA-CoA formation at the expense of PG formation when 5 microM AA was used as substrate concentration. The experiments utilizing Fe(2+) and an electron spin resonance (ESR) revealed that 15-HPEPE elicits these effects in the form of hydroperoxy adduct. These results suggest that 15-HPEPE, but not EPA, has the potential to shift AA away from COX pathway into ACS pathway at low substrate concentration (close to the physiological concentration of AA).  相似文献   

19.
Components I and II of cytochrome cd1 which had different spectral features were purified from the aerobic photosynthetic bacterium, Erythrobacter sp. strain OCh 114. Component I showed an absorption maxima at 700 and 406 nm in the oxidized form, and at 621, 552.5, 548 and 416 nm in the reduced form. Component II showed an absorption maxima at 635 and 410 nm in the oxidized form and at 628, 552.5, 548 and 417 nm in the reduced form. The relative molecular mass, Mr, of both cytochromes was determined to be 135,000 with two identical subunits. Components I and II showed pI values of 7.6 and 6.8, respectively. The redox potential of hemes ranged from +234 mV to +242 mV, except for the heme d1 of component I (Em7 = +134 mV). Components I and II showed both cytochrome c oxidase and nitrite reductase activities. Cytochrome c oxidase activity was strongly inhibited by a low concentration of nitrite and cyanide. Erythrobacter cytochromes c-551 and c-552 were utilized as electron donors for the cytochrome c oxidase reaction. The high affinity of cytochrome c-552 to component II (Km = 1.27 microM) suggested a physiological significance for this cytochrome. Erythrobacter cytochromes cd1 are unique in their presence in cells grown under aerobic conditions as compared to other bacterial cytochromes cd1 which are formed only under denitrifying conditions.  相似文献   

20.
Two types of voltage-dependent Ca(2+) channels have been identified in heart: high (I(CaL)) and low (I(CaT)) voltage-activated Ca(2+) channels. In guinea pig ventricular myocytes, low voltage-activated inward current consists of I(CaT) and a tetrodotoxin (TTX)-sensitive I(Ca) component (I(Ca(TTX))). In this study, we reexamined the nature of low-threshold I(Ca) in dog atrium, as well as whether it is affected by Na(+) channel toxins. Ca(2+) currents were recorded using the whole-cell patch clamp technique. In the absence of external Na(+), a transient inward current activated near -50 mV, peaked at -30 mV, and reversed around +40 mV (HP = -90 mV). It was unaffected by 30 microM TTX or micromolar concentrations of external Na(+), but was inhibited by 50 microM Ni(2+) (by approximately 90%) or 5 microM mibefradil (by approximately 50%), consistent with the reported properties of I(CaT). Addition of 30 microM TTX in the presence of Ni(2+) increased the current approximately fourfold (41% of control), and shifted the dose-response curve of Ni(2+) block to the right (IC(50) from 7.6 to 30 microM). Saxitoxin (STX) at 1 microM abolished the current left in 50 microM Ni(2+). In the absence of Ni(2+), STX potently blocked I(CaT) (EC(50) = 185 nM) and modestly reduced I(CaL) (EC(50) = 1.6 microM). While TTX produced no direct effect on I(CaT) elicited by expression of hCa(V)3.1 and hCa(V)3.2 in HEK-293 cells, it significantly attenuated the block of this current by Ni(2+) (IC(50) increased to 550 microM Ni(2+) for Ca(V)3.1 and 15 microM Ni(2+) for Ca(V)3.2); in contrast, 30 microM TTX directly inhibited hCa(V)3.3-induced I(CaT) and the addition of 750 microM Ni(2+) to the TTX-containing medium led to greater block of the current that was not significantly different than that produced by Ni(2+) alone. 1 microM STX directly inhibited Ca(V)3.1-, Ca(V)3.2-, and Ca(V)3.3-mediated I(CaT) but did not enhance the ability of Ni(2+) to block these currents. These findings provide important new implications for our understanding of structure-function relationships of I(CaT) in heart, and further extend the hypothesis of a parallel evolution of Na(+) and Ca(2+) channels from an ancestor with common structural motifs.  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