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.
Mechanism of oxygen activation by tyrosine hydroxylase   总被引:5,自引:0,他引:5  
T A Dix  D M Kuhn  S J Benkovic 《Biochemistry》1987,26(12):3354-3361
The mechanism by which the tetrahydropterin-requiring enzyme tyrosine hydroxylase (TH) activates dioxygen for substrate hydroxylation was explored. TH contains one ferrous iron per subunit and catalyzes the conversion of its tetrahydropterin cofactor to a 4a-carbinolamine concomitant with substrate hydroxylation. These results are in accord with shared mechanisms of oxygen activation by TH and the more commonly studied tetrahydropterin-dependent enzyme phenylalanine hydroxylase (PAH) and strongly suggest that a peroxytetrahydropterin is the hydroxylating species generated during TH turnover. In addition, TH can also utilize H2O2 as a cofactor for substrate hydroxylation, a result not previously established for PAH. A detailed mechanism for the reaction is proposed. While the overall pattern of tetrahydropterin-dependent oxygen activation by TH and PAH is similar, the H2O2-dependent hydroxylation performed by TH provides an indication that subtle differences in the Fe ligand field exist between the two enzymes. The mechanistic ramifications of these results are briefly discussed.  相似文献   

2.
Toluate dioxygenase (TADO) of Pseudomonas putida mt-2 catalyzes the dihydroxylation of a broad range of substituted benzoates. The two components of this enzyme were hyperexpressed and anaerobically purified. Reconstituted TADO had a specific activity of 3.8 U/mg with m-toluate, and each component had a full complement of their respective Fe(2)S(2) centers. Steady-state kinetics data obtained by using an oxygraph assay and by varying the toluate and dioxygen concentrations were analyzed by a compulsory order ternary complex mechanism. TADO had greatest specificity for m-toluate, displaying apparent parameters of KmA = 9 +/- 1 microM, k(cat) = 3.9 +/- 0.2 s(-1), and K(m)O(2) = 16 +/- 2 microM (100 mM sodium phosphate, pH 7.0; 25 degrees C), where K(m)O(2) represents the K(m) for O(2) and KmA represents the K(m) for the aromatic substrate. The enzyme utilized benzoates in the following order of specificity: m-toluate > benzoate approximately 3-chlorobenzoate > p-toluate approximately 4-chlorobenzoate > o-toluate approximately 2-chlorobenzoate. The transformation of each of the first five compounds was well coupled to O(2) utilization and yielded the corresponding 1,2-cis-dihydrodiol. In contrast, the transformation of ortho-substituted benzoates was poorly coupled to O(2) utilization, with >10 times more O(2) being consumed than benzoate. However, the apparent K(m) of TADO for these benzoates was >100 microM, indicating that they do not effectively inhibit the turnover of good substrates.  相似文献   

3.
Human tyrosine 3-monooxygenase (tyrosine hydroxylase) exists as four different isozymes (TH1-TH4), generated by alternative splicing of pre-mRNA. Recombinant TH1, TH2 and TH4 were expressed in high yield in Escherichia coli. The purified isozymes revealed high catalytic activity [when reconstituted with Fe(II)] and stability at neutral pH. The isozymes as isolated contained 0.04-0.1 atom iron and 0.02-0.06 atom zinc/enzyme subunit. All three isozymes were rapidly activated (13-40-fold) by incubation with Fe(II) salts (concentration of iron at half-maximal activation = 6-14 microM), and were inhibited by other divalent metal ions, e.g. Zn(II), Co(II) and Ni(II). They all bind stoichiometric amounts of Fe(II) and Zn(II) with high affinity (Kd = 0.2-3 microM at pH 5.4-6.5). Similar time courses were observed for binding of Fe(II) and enzyme activation. In the absence of any free Fe(II) or Zn(II), the metal ions were released from the reconstituted isozymes. The dissociation was favoured by acidic pH, as well as by the presence of metal chelators and dithiothreitol. The potency of metal chelators to remove iron from the hydroxylase correlated with their ability to inhibit the enzyme activity. These studies show that tyrosine hydroxylase binds iron reversibly and that its catalytic activity is strictly dependent on the presence of this metal.  相似文献   

4.
Escherichia coli expresses an inducible flavohemoglobin possessing robust NO dioxygenase activity. At 37 degrees C, the enzyme shows a maximal turnover number (V(max)) of 670 s(-1) and K(m) values for NADH, NO, and O(2) equal to 4.8, 0.28, and approximately 100 microM, respectively. Individual reduction, ligand binding, and NO dioxygenation reactions were examined at 20 degrees C, where V(max) is approximately 94 s(-1). Reduction by NADH occurs in two steps. NADH reduces bound FAD with a rate constant of approximately 15 microM(-1) s(-1), and heme iron is reduced by FADH(2) with a rate constant of 150 s(-1). Dioxygen binds tightly to reduced flavohemoglobin, with association and dissociation rate constants equal to 38 microM(-1) s(-1) and 0.44 s(-1), respectively, and the oxygenated flavohemoglobin dioxygenates NO to form nitrate. NO also binds reversibly to reduced flavohemoglobin in competition with O(2), dissociates slowly, and inhibits NO dioxygenase activity at [NO]/[O(2)] ratios of 1:100. Replacement of the heme pocket B10 tyrosine with phenylalanine increases the O(2) dissociation rate constant approximately 80-fold and reduces NO dioxygenase activity approximately 30-fold, demonstrating the importance of the tyrosine hydroxyl for O(2) affinity and NO scavenging activity. At 37 degrees C, V(max)/K(m)(NO) is 2,400 microM(-1) s(-1), demonstrating that the enzyme is extremely efficient at converting toxic NO into nitrate under physiological conditions.  相似文献   

5.
The Na+,K(+)-ATPase is a membrane-bound, sulfhydryl-containing protein whose activity is critical to maintenance of cell viability. The susceptibility of the enzyme to radical-induced membrane lipid peroxidation was determined following incorporation of a purified Na+,K(+)-ATPase into soybean phosphatidylcholine liposomes. Treatment of liposomes with Fenton's reagent (Fe2+/H2O2) resulted in malondialdehyde formation and total loss of Na+,K(+)-ATPase activity. At 150 microM Fe2+/75 microM H2O2, vitamin E (5 mol%) totally prevented lipid peroxidation but not the loss of enzyme activity. Lipid peroxidation initiated by 25 microM Fe2+/12.5 microM H2O2 led to a loss of Na+,K(+)-ATPase activity, however, vitamin E (1.2 mol%) prevented both malondialdehyde formation and loss of enzyme activity. In the absence of liposomes, there was complete loss of Na+,K(+)-ATPase activity in the presence of 150 microM Fe2+/75 microM H2O2, but little effect by 25 microM Fe2+/12.5 microM H2O2. The activity of the enzyme was also highly sensitive to radicals generated by the reaction of Fe2+ with cumene hydroperoxide, t-butylhydroperoxide, and linoleic acid hydroperoxide. Lipid peroxidation initiated by 150 microM Fe2+/150 microM Fe3+, an oxidant which may be generated by the Fenton's reaction, inactivated the enzyme. In this system, inhibition of malondialdehyde formation by vitamin E prevented loss of Na+,K(+)-ATPase activity. These data demonstrate the susceptibility of the Na+,K(+)-ATPase to radicals produced during lipid peroxidation and indicate that the ability of vitamin E to prevent loss of enzyme activity is highly dependent upon both the nature and the concentration of the initiating and propagating radical species.  相似文献   

6.
Previous results from our laboratory have shown that NADH-supported electron flow through the Escherichia coli respiratory chain promotes the reduction of cupric ions to Cu(I), which mediates damage of the respiratory system by hydroperoxides. The aim of this work was to characterize the NADH-linked cupric reductase activity from the E. coli respiratory chain. We have used E. coli strains that either overexpress or are deficient in the NADH dehydrogenase-2 (NDH-2) to demonstrate that this membrane-bound protein catalyzes the electron transfer from NADH to Cu(II), but not to Fe(III). We also show that purified NDH-2 exhibits NADH-supported Cu(II) reductase activity in the presence of either FAD or quinone, but is unable to reduce Fe(III). The K(m) values for free Cu(II) were 32 +/- 5 pM in the presence of saturating duroquinone and 22 +/- 2 pM in the presence of saturating FAD. The K(m) values for NADH were 6.9 +/- 1.5 microM and 6.1 +/- 0.7 microM in the presence of duroquinone and FAD, respectively. The quinone-dependent Cu(II) reduction occurred through both O(*-)(2)-mediated and O(*-)(2)-independent pathways, as evidenced by the partial inhibitory effect (30-50%) of superoxide dismutase, by the reaction stoichiometry, and by the enzyme turnover numbers for NADH and Cu(II). The cupric reductase activity of NDH-2 was dependent on thiol groups which were accessible to p-chloromercuribenzoate at low, but not at high, ionic strength of the medium, a fact apparently connected to a conformational change of the protein. To our knowledge, this is the first protein with cupric reductase activity to be isolated and characterized in its biochemical properties.  相似文献   

7.
An early development-specific soluble 55 kDa Ca(2+)-dependent protein kinase has been purified to homogeneity from sandalwood somatic embryos and biochemically characterized. The purified enzyme, swCDPK, resolved into a single band on 10% polyacrylamide gels, both under denaturing and non-denaturing conditions. swCDPK activity was strictly dependent on Ca(2+), K(0.5) (apparent binding constant) for Ca(2+)-activation of substrate phosphorylation activity being 0.7 microM and for autophosphorylation activity approximately 50 nM. Ca(2+)-dependence for activation, CaM-independence, inhibition by CaM-antagonist (IC(50) for W7=6 microM, for W5=46 microM) and cross-reaction with polyclonal antibodies directed against the CaM-like domain of soybean CDPK, confirmed the presence of an endogenous CaM-like domain in the purified enzyme. Kinetic studies revealed a K(m) value of 1.3 mg/ml for histone III-S and a V(max) value of 0.1 nmol min(-1) mg(-1). The enzyme exhibited high specificity for ATP with a K(m) value of 10 nM. Titration with calcium resulted in the enhancement of intrinsic emission fluorescence of swCDPK and a shift in the lambda(max) emission from tryptophan residues. A reduction in the efficiency of non-radiative energy transfer from tyrosine to tryptophan residues was also observed. These are taken as evidence for the occurrence of Ca(2+)-induced conformational change in swCDPK. The emission spectral properties of swCDPK in conjunction with Ca(2+) levels required for autophosphorylation and substrate phosphorylation help understand mode of Ca(2+) activation of this enzyme.  相似文献   

8.
Mills SA  Goto Y  Su Q  Plastino J  Klinman JP 《Biochemistry》2002,41(34):10577-10584
A recent report by Mills and Klinman [Mills, S. A., and Klinman, J. P. (2000) J. Am. Chem. Soc. 122, 9897-9904] described the preparation and initial characterization of a cobalt-substituted form of the copper amine oxidase from Hansenula polymorpha (HPAO). This enzyme was found to be fully catalytically active at saturating substrate concentrations, but with a K(m) for O(2) approximately 70-fold higher than that of the copper-containing, wild-type enzyme. Herein, we report a detailed analysis of the mechanism of catalysis for the wild-type and the cobalt-substituted forms of HPAO. Both forms of enzyme are concluded to utilize the same mechanism for oxygen reduction, involving initial, rate-limiting electron transfer from the reduced cofactor of the enzyme to prebound dioxygen. Superoxide formed in this manner is stabilized by the active site metal, facilitating the transfer of a second electron and two protons to form the product hydrogen peroxide. The elevated K(m) for O(2) at the dioxygen binding site in Co-substituted HPAO, relative to that of wild-type HPAO, is proposed to be due to a change in the net charge at the adjacent metal site from +1 (cupric hydroxide) in wild-type enzyme to +2 (cobaltous H(2)O) in cobalt-substituted HPAO.  相似文献   

9.
Summary. We have investigated the idea that nicotinamide, a non-selective inhibitor of the sentinel enzyme Poly(ADP-ribose) polymerase-I (PARP-1), provides neuroprotection against the long-term neurological changes induced by perinatal asphyxia. Perinatal asphyxia was induced in vivo by immersing foetuses-containing uterine horns removed from ready-to-deliver rats into a water bath for 20 min. Sibling caesarean-delivered pups were used as controls. The effect of perinatal asphyxia on neurocircuitry development was studied in vitro with organotypic cultures from substantia nigra, neostriatum and neocortex, platted on a coverslip 3 days after birth. After approximately one month in vitro (DIV 25), the cultures were treated for immunocytochemistry to characterise neuronal phenotype with markers against the N-methyl-D-aspartate receptor subunit 1 (NR1), the dopamine pacemaker enzyme tyrosine hydroxylase (TH), and nitric oxide synthase (NOS), the enzyme regulating the bioavailability of NO. Nicotinamide (0.8 mmol/kg, i.p.) or saline was administered to asphyctic and caesarean-delivered pups 24, 48 and 72 h after birth. It was found that nicotinamide treatment prevented the effect of perinatal asphyxia on several neuronal parameters, including TH- and NOS-positive neurite atrophy and NOS-positive neuronal loss; supporting the idea that nicotinamide constitutes a therapeutic alternative for the effects produced by sustained energy-failure conditions, as occurring during perinatal asphyxia.  相似文献   

10.
11.
Involvement of Ca2+/calmodulin-dependent protein kinase II (Ca2+/CaM-kinase II) on the phosphorylation of tyrosine hydroxylase (TH, EC.1.14.16.2) in rat pheochromocytoma, PC12h cells was examined using KN-62, 1-[N,O-Bis(5-isoquinolinsulfonyl)-N-methyl-L-tyrosyl]-4-phenylpipe razine, a selective inhibitor of Ca2+/CaM-kinase II. Both the enhanced phosphorylation of TH and the activated L-3,4-dihydroxyphenylalanine (DOPA) formation in the high K+ depolarization were inhibited by 10 microM KN-62. After incubation of PC12h cells with 10 microM KN-62 for 1 hr, the activation of TH with 3 min incubation of 56 mM K+ was reduced to the basal activity. However, KN-62 did not directly affect the activity of purified rat TH at pH 6.0 or 7.0. These results indicate that Ca2+/CaM-kinase II phosphorylates and activates TH of PC12h cells in the high K+ depolarization.  相似文献   

12.
Azoreductase, an enzyme catalyzing the reductive cleavage of the azo bond of methyl red (MR) and related dyes, was purified to electrophoretic homogeneity from Enterobacter agglomerans. This bacterial strain, isolated from dye-contaminated sludge, has a higher ability to grow, under aerobic conditions, on culture medium containing 100mg/L of MR. The enzyme was purified approximately 90-fold with 20% yield by ammonium sulfate precipitation, followed by three steps of column chromatography (gel-filtration, anion-exchange, and dye-affinity). The purified enzyme is a monomer with a molecular weight of 28,000 Da. The maximal azoreductase activity was observed at pH 7.0 and at 35 degrees C. This activity was NADH dependent. The K(m) values for both NADH and MR were 58.9 and 29.4 microM, respectively. The maximal velocity (V(max)) was 9.2 micromol of NADH min(-1)mg(-1). The purified enzyme is inhibited by several metal ions including Fe(2+) and Cd(2+).  相似文献   

13.
Schwartz B  Dove JE  Klinman JP 《Biochemistry》2000,39(13):3699-3707
A detailed kinetic analysis of oxygen consumption during TPQ biogenesis has been carried out on a yeast copper amine oxidase. O(2) is consumed in a single, exponential phase, the rate of which responds linearly to dissolved oxygen concentration. This behavior is observed up to conditions of maximally obtainable oxygen concentrations. In contrast, no viscosity effect is observed on rate, implicating a high K(m) for O(2). Binding of oxygen appears to occur faster than its consumption and to result in displacement of the precursor tyrosine onto copper to form a charge-transfer species, described in the the preceding paper of this issue [Dove, J. E., Schwartz, B., Williams, N. K., and Klinman, J. P. (2000) Biochemistry 39, 3690-3698). Reaction between this intermediate and O(2) is proposed to occur in a rate-limiting step, and to proceed more rapidly when the tyrosine is deprotonated. This rate-limiting step in cofactor biogenesis does not display a solvent isotope effect and is, thus, uncoupled from proton transfer. Comparisons are drawn between the proposed biogenesis mechanism and that for the oxidation of reduced cofactor during catalytic turnover in the mature enzyme.  相似文献   

14.
Enzymatically active mouse tyrosine hydroxylase (TH) was successfully expressed at a high level in Escherichia coli using a T7 RNA polymerase directed expression system. The specific activity of mouse TH in E. coli cell lysate was 7.5 nmol/mg protein/min. Kinetic characteristics of recombinant TH were examined. Km for tyrosine and (6R)-tetrahydrobiopterin (6R-BH4) cofactor were determined to be 7.2 microM (420 microM 6R-BH4), 19 microM [( 6R-BH4] less than 55 microM, 20 microM tyrosine) and 54 microM [( 6R-BH4] greater than 55 microM, 20 microM tyrosine), respectively. These were in good agreement with previously reported values for this enzyme.  相似文献   

15.
The 3-hydroxypropionate cycle is a new autotrophic CO(2) fixation pathway in Chloroflexus aurantiacus and some archaebacteria. The initial step is acetyl-coenzyme A (CoA) carboxylation to malonyl-CoA by acetyl-CoA carboxylase, followed by NADPH-dependent reduction of malonyl-CoA to 3-hydroxypropionate. This reduction step was studied in Chloroflexus aurantiacus. A new enzyme was purified, malonyl-CoA reductase, which catalyzed the two-step reduction malonyl-CoA + NADPH + H(+) --> malonate semialdehyde + NADP(+) + CoA and malonate semialdehyde + NADPH + H(+) --> 3-hydroxypropionate + NADP(+). The bifunctional enzyme (aldehyde dehydrogenase and alcohol dehydrogenase) had a native molecular mass of 300 kDa and consisted of a single large subunit of 145 kDa, suggesting an alpha(2) composition. The N-terminal amino acid sequence was determined, and the incomplete gene was identified in the genome database. Obviously, the enzyme consists of an N-terminal short-chain alcohol dehydrogenase domain and a C-terminal aldehyde dehydrogenase domain. No indication of the presence of a prosthetic group was obtained; Mg(2+) and Fe(2+) stimulated and EDTA inhibited activity. The enzyme was highly specific for its substrates, with apparent K(m) values of 30 microM malonyl-CoA and 25 microM NADPH and a turnover number of 25 s(-1) subunit(-1). The specific activity in autotrophically grown cells was 0.08 micromol of malonyl-CoA reduced min(-1) (mg of protein)(-1), compared to 0.03 micromol min(-1) (mg of protein)(-1) in heterotrophically grown cells, indicating downregulation under heterotrophic conditions. Malonyl-CoA reductase is not required in any other known pathway and therefore can be taken as a characteristic enzyme of the 3-hydroxypropionate cycle. Furthermore, the enzyme may be useful for production of 3-hydroxypropionate and for a coupled spectrophotometric assay for activity screening of acetyl-CoA carboxylase, a target enzyme of potent herbicides.  相似文献   

16.
The aim of this work was to study the hydroxylation of N, N-dimethyltyramine (DMTA) by tyrosinase in the presence of hydrogen peroxide, a reaction that does not take place without the addition of the hydrogen peroxide. Some properties of this hydroxylating activity are analyzed. The kinetic parameters of mushroom tyrosinase toward hydrogen peroxide (K(m) = 0.5 mM, V(m) = 11 microM/min, V(m)/K(m) = 2.2 x 10(-2) min(-1)) and toward DMTA (K(m) = 0.3 mM, V(m) = 4.8 microM/min, V(m)/K(m) = 16 x 10(-2) min(-1)) were evaluated. There was a lag period, which was similar to the characteristic lag of monophenolase activity at the expense of molecular oxygen. The length of this lag phase decreased with increasing hydrogen peroxide concentration, and disappeared at approximately 0.5 mM H(2)O(2). However, the lag was longer with higher DMTA concentrations. The pH optimum range for this hydroxylating activity was 6.0 to 7.0. The lag also varied with pH, increasing at pH values higher than 6.7. The presence of hydrogen peroxide is necessary for the oxidation of DMTA, as is the presence of active enzyme since the reaction was completely inhibited when selective tyrosinase inhibitors were added.  相似文献   

17.
Arachidonate 5-lipoxygenase of a 10,000 x g supernatant from RBL-1 cell homogenate was studied by a continuous assay measuring enzyme-catalysed oxygen consumption. Parallel HPLC and TLC analysis of arachidonic acid metabolites revealed that the oxygen consumption measured is solely due to 5-lipoxygenation of arachidonic acid. Oxygen consumption by this lipoxygenase was strictly dependent upon Ca2+, ATP and 5-HPETE. Removal of any of these three cofactors caused a complete inhibition of enzyme activity. Addition of the missing cofactor instantly restored the 5-lipoxygenase-dependent consumption of oxygen which remained linear for 10-20 s. Later on the velocity of the reaction decreased and after 2-3 min the enzyme became inactivated. Kinetic data were obtained from the initial velocity of the reaction using constant and saturating concentrations of CaCl2 and ATP. From Lineweaver-Burk plots substrate inhibition is evident for arachidonic acid concentrations greater than 45-50 microM. Km(app) for arachidonic acid is 182 +/- 16 microM (mean +/- SD, n = 5) and Vmax(app) is 425 +/- 140 nmol O2/(min x mg protein) (mean +/- SD, n = 5).  相似文献   

18.
The final step in the biosynthesis of the plant hormone ethylene is catalyzed by the non-heme iron-containing enzyme 1-aminocyclopropane-1-carboxylic acid (ACC) oxidase (ACCO). ACC is oxidized at the expense of O(2) to yield ethylene, HCN, CO(2), and two waters. Continuous turnover of ACCO requires the presence of ascorbate and HCO(3)(-) (or an alternative form), but the roles played by these reagents, the order of substrate addition, and the mechanism of oxygen activation are controversial. Here these issues are addressed by development of the first functional single turnover system for ACCO. It is shown that 0.35 mol ethylene/mol Fe(II)ACCO is produced when the enzyme is combined with ACC and O(2) in the presence of HCO(3)(-) but in the absence of ascorbate. Thus, ascorbate is not required for O(2) activation or product formation. Little product is observed in the absence of HCO(3)(-), demonstrating the essential role of this reagent. By monitoring the EPR spectrum of the sample during single turnover, it is shown that the active site Fe(II) oxidizes to Fe(III) during the single turnover. This suggests that the electrons needed for catalysis can be derived from a fraction of the initial Fe(II)ACCO instead of ascorbate. Addition of ascorbate at 10% of its K(m) value significantly accelerates both iron oxidation and ethylene formation, suggesting a novel high-affinity effector role for this reagent. This role can be partially mimicked by a non-redox-active ascorbate analog. A mechanism is proposed that begins with ACC and O(2) binding, iron oxidation, and one-electron reduction to form a peroxy intermediate. Breakdown of this intermediate, perhaps by HCO(3)(-)-mediated proton transfer, is proposed to yield a high-valent iron species, which is the true oxidizing reagent for the bound ACC.  相似文献   

19.
A nitrophenol oxygenase which stoichiometrically converted ortho-nitrophenol (ONP) to catechol and nitrite was isolated from Pseudomonas putida B2 and purified. The substrate specificity of the enzyme was broad and included several halogen- and alkyl-substituted ONPs. The oxygenase consisted of a single polypeptide chain with a molecular weight of 58,000 (determined by gel filtration) or 65,000 (determined on a sodium dodecyl sulfate-polyacrylamide gel). The enzymatic reaction was NADPH dependent, and one molecule of oxygen was consumed per molecule of ONP converted. Enzymatic activity was stimulated by magnesium or manganese ions, whereas the addition of flavin adenine dinucleotide, flavin mononucleotide, or reducing agents had no effect. The apparent Kms for ONP and NADPH were 8 and 140 microM, respectively. 2,4-Dinitrophenol competitively (Ki = 0.5 microM) inhibited ONP turnover. The optimal pH for enzyme stability and activity was in the range of 7.5 to 8.0. At 40 degrees C, the enzyme was totally inactivated within 2 min; however, in the presence of 1 mM ONP, 40% of the activity was recovered, even after 10 min. Enzymatic activity was best preserved at -20 degrees C in the presence of 50% glycerol.  相似文献   

20.
Biphasic electrical field stimulation (0.5-5 Hz, 2 ms, 25 V, 3 min) and high K+ (10-30 mM, 5 min) released endogenous 3,4-dihydroxyphenylalanine (DOPA) from superfused rat striatal slices. Characteristics of the DOPA release were compared with those of 3,4-dihydroxyphenylethylamine (dopamine, DA). Electrical stimulation at 2 Hz evoked DOPA and DA over similar time courses. alpha-Methyl-p-tyrosine (0.2 mM) markedly reduced release of DOPA but not of DA. Maximal release (0.3 pmol) of DOPA was obtained at 2 Hz and at 15 mM K+. The impulse-evoked release of DOPA and DA was completely tetrodotoxin (0.3 microM) sensitive and Ca2+ dependent and the 15 mM K+-evoked release was also Ca2+ dependent. On L-[3,5-3H]tyrosine (1 microM) superfusion, high K+ (15 and 60 mM) released DOPA and DA together with concentration-dependent decreases in tyrosine 3-monooxygenase (EC 1.14.16.2) activity as indicated by [3H]H2O formation, followed by concentration-dependent increases after DOPA and DA release ended. These findings suggest that striatal DOPA is released by a Ca2+-dependent excitation-secretion coupling process similar to that involved in transmitter release.  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