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 15 毫秒
1.
Hyoscyamine 6 beta-hydroxylase, a 2-oxoglutarate-dependent dioxygenase that catalyzes the hydroxylation of l-hyoscyamine to 6 beta-hydroxyhyoscyamine in the biosynthetic pathway leading to scopolamine [Hashimoto, T. & Yamada, Y. (1986) Plant Physiol. 81, 619-625] was purified 310-fold from root cultures of Hyoscyamus niger L. The enzyme has an average Mr of 41,000 as determined by gel filtration on Superose 12 and exhibited maximum activity at pH 7.8 l-Hyoscyamine and 2-oxoglutarate are required for the enzyme activity, with respective Km values of 35 microM and 43 microM. Fe2+, catalase and a reductant such as ascorbate significantly activated the enzyme. 2-Oxoglutarate was not replaced by any of ten other oxo acids tested, nor was Fe2+ by nine other divalent cations tested. The enzyme was inhibited moderately by EDTA, Tiron and various oxo acids and aliphatic dicarboxylic acids, and strongly by nitroblue tetrazolium and divalent cations Mn2+, Co2+, Ni2+, Cu2+, Zn2+, Cd2+ and Hg2+. Several pyridine dicarboxylates and o-dihydroxyphenyl derivatives inhibited the hydroxylase. Pyridine 2,4-dicarboxylate and 3,4-dihydroxybenzoate are competitive inhibitors with respect to 2-oxoglutarate with the respective Ki values of 9 microM and 90 microM. Several alkaloids with structures similar to l-hyoscyamine were hydroxylated by the enzyme at the C-6 position of the tropane moiety. The enzyme preparation also epoxidized 6,7-dehydrohyoscyamine, a hypothetical precursor of scopolamine, to scopolamine (Km 10 microM). This epoxidation reaction required the same co-factors as the hydroxylation reaction and the epoxidase activities were found in the same fractions with the hydroxylase activities during purification. Two possible pathways for scopolamine biosynthesis are discussed in the light of the hydroxylase and epoxidase activities found in the partially purified preparation of hyoscyamine 6 beta-hydroxylase.  相似文献   

2.
Abstract: The substrate dependence of tryptophan hydroxylase activity in rat striatal synaptosomes was examined. The K m for tryptophan in air is 8-13 μM, comparable to the concentration present in cerebrospinal fluid. The reaction is inhibited by amino acids with large nonpolar side chains. For leucine the inhibition appears competitive; it results from a decrease in the steady state levels of intrasynaptosomal tryptophan. The oxygen Km at 10 μM-tryptophan is 3-4 mm Hg, which is increased when the reaction is assayed with low levels of tryptophan in the medium, and in the presence of amino acids such as leucine. Similarly, the tryptophan K m is increased at low oxygen tensions; at 9 mm Hg of oxygen it is 23 μM. These interactions between tryptophan and oxygen dependence of the reaction are discussed in terms of likely physiological significance and implications for the pharmacological use of tryptophan.  相似文献   

3.
The enzyme tyrosine hydroxylase catalyzes the first step in the biosynthesis of dopamine, norepinephrine, and epinephrine. Tyrosine hydroxylase is a substrate for cyclic AMP-dependent protein kinase as well as other protein kinases. We determined the Km and Vmax of rat pheochromocytoma tyrosine hydroxylase for cyclic AMP-dependent protein kinase and obtained values of 136 microM and 7.1 mumol/min/mg of catalytic subunit, respectively. These values were not appreciably affected by the substrates for tyrosine hydroxylase (tyrosine and tetrahydrobiopterin) or by feedback inhibitors (dopamine and norepinephrine). The high Km of tyrosine hydroxylase correlates with the high content of tyrosine hydroxylase in catecholaminergic cells. We also determined the kinetic constants for peptides modeled after actual or potential tyrosine hydroxylase phosphorylation sites. We found that the best substrates for cyclic AMP-dependent protein kinase were those peptides corresponding to serine 40. Tyrosine hydroxylase (36-46), for example, exhibited a Km of 108 microM and a Vmax of 6.93 mumol/min/mg of catalytic subunit. The next best substrate was the peptide corresponding to serine 153. The peptide containing the sequence conforming to serine 19 was a very poor substrate, and that conforming to serine 172 was not phosphorylated to any significant extent. The primary structure of the actual or potential phosphorylation sites is sufficient to explain the substrate behavior of the native enzyme.  相似文献   

4.
A reproducible purification procedure of native tyrosine hydroxylase (L-tyrosine, tetrahydropteridine : oxygen oxidoreductase (3-hydroxylating), EC 1.14.16.2) from the soluble fraction of the bovine adrenal medulla has been established. This procedure accomplished a 90-fold purification with a recovery of 30% of the activity. This purified enzyme served for studying the kinetic properties of tyrosine hydroxylase using (6R)-L-erythro-1',2'-dihydroxypropyltetrahydropterin [(6R)-L-erythro-tetrahydrobiopterin] as cofactor, which is supposed to be a natural cofactor. Two different Km values for tyrosine, oxygen and natural (6R)-L-erythro-tetrahydrobiopterin itself were obtained depending on the concentration of the tetrahydrobiopterin cofactor. In contrast, when unnatural (6S)-L-erythro-tetrahydrobiopterin was used as cofactor, a single Km value for each tyrosine, oxygen and the cofactor was obtained independent of the cofactor concentration. The lower Km value for (6R)-L-erythro-tetrahydrobiopterin was close to the tetrahydrobiopterin concentration in tissue, indicating a high affinity of the enzyme to the natural cofactor under the in vivo conditions. Tyrosine was inhibitory at 100 microM with (6R)-L-erythro-tetrahydrobiopterin as cofactor, and the inhibition by tyrosine was dependent on the concentrations of both pterin cofactor and oxygen. Oxygen at concentrations higher than 4.8% was also inhibitory with (6R)-L-erythro-tetrahydrobiopterin as cofactor.  相似文献   

5.
Pheochromocytoma tyrosine hydroxylase was reported to have unusual catalytic properties, which might be unique to the tumor enzyme (Dix, T. A., Kuhn, D. M., and Benkovic, S. J. (1987) Biochemistry 24, 3354-3361). Two such properties, namely the apparent inability to hydroxylate phenylalanine and an unprecedented reactivity with hydrogen peroxide were investigated further in the present study. Tyrosine hydroxylase was purified to apparent homogeneity from cultured pheochromocytoma PC12 cells. The purified tumor enzyme was entirely dependent on tetrahydrobiopterin (BH4) for the hydroxylation of tyrosine to 3,4-dihydroxyphenylalanine and hydrogen peroxide could not substitute for the natural cofactor. Indeed, in the presence of BH4, increasing concentrations of hydrogen peroxide completely inhibited enzyme activity. The PC12 hydroxylase exhibited typical kinetics of tyrosine hydroxylation exhibited typical kinetics of tyrosine hydroxylation, both as a function of tyrosine (S0.5 Tyr = 15 microM) and BH4 (apparent Km BH4 = 210 microM). In addition, the enzyme catalyzed the hydroxylation of substantial amounts of phenylalanine to tyrosine and 3,4-dihydroxyphenylalanine (apparent Km Phe = 100 microM). Phenylalanine did not inhibit the enzyme in the concentrations tested, whereas tyrosine showed typical substrate inhibition at concentrations greater than or equal to 50 microM. At higher substrate concentrations, the rate of phenylalanine hydroxylation was equal to or exceeded that of tyrosine. Essentially identical results were obtained with purified tyrosine hydroxylase from pheochromocytoma PC18 cells. The data suggest that the tumor enzyme has the same substrate specificity and sensitivity to hydrogen peroxide as tyrosine hydroxylase from other tissues.  相似文献   

6.
Tyrosine hydroxylation was studied in intact cells of mouse neuroblastoma clone N1E-115 which have high levels of tyrosine 3-monooxygenase (EC 1.14.16.2) and which have been fully characterized for tyrosine transport. Measurement of [3H]OH formed from L-[3,5(-3)H]tyrosine in the medium was the method of assay and [3H]OH formed was stoichiometric with the formation of L-[3H]3,4-dihydroxyphenylalanine. Tyrosine hydroxylation was dependent on time of incubation, cell number, and the concentration of [3H]tyrosine in the medium. From velocity vs. [3H]tyrosine concentration experiments, two apparent Km values were obtained: Km1 = 10 +/- 2 microM; Km2 = 140 +/- 10 microM. Substrate inhibition occurred with tyrosine concentrations between 20 and 50 microM. The reaction was twice as fast at pH 5.5 as at pH 7.4. alpha,alpha'-Dipyridyl (1 mM) caused major inhibition (75%) when [3H]tyrosine concentration was 10 microM. L-3-Iodotyrosine was a competitive inhibitor with Ki = 0.3 microM. Dopamine was a non-competitive inhibitor with Ki = 500 microM. 1-Norepinephrine had no effect. These results show that the hydroxylation of tyrosine by living N1E-115 cells has many of the properties of the reaction catalyzed by purified tyrosine 3-monooxygenase from normal tissue.  相似文献   

7.
Tyrosine hydroxylase phosphatase activity in rat caudate nucleus was separated into three peaks by chromatography on DEAE-cellulose. [32P]Tyrosine hydroxylase phosphorylated by cyclic AMP-dependent protein kinase was dephosphorylated only by the major peak eluting at 0.3 M NaCl, while tyrosine hydroxylase phosphorylated by Ca2+-calmodulin-dependent protein kinase was also dephosphorylated by two calcium-inhibited phosphatases. The Vmax of the enzyme in the major DEAE peak was increased by 10 microM tetrahydrobiopterin (BH4) from 0.78 to 5.0 fmol min-1 mg-1 while the Km was only slightly affected, increasing from 45 to 62 pM. The activation could not be reversed by dilution. On Sephadex G-200, the enzyme was found to consist of two major forms with molecular masses of 420 and 100 kDa. In contrast to the activation of liver phosphatases by freezing with beta-mercaptoethanol, activation by tetrahydrobiopterin was not associated with a shift in the molecular weight of the phosphatase to lower molecular weight forms. Other reduced pterins, including tetrahydroneopterin, 6-methyltetrahydropterin, and 5-methyltetrahydrofolate, also activated the enzyme, while oxidized pterins had no effect. GTP, the metabolic precursor of tetrahydrobiopterin, was a potent inhibitor of the phosphatase reaction, inhibiting by 65% at a concentration of 1 microM. These findings suggest a close regulatory interrelationship between the tetrahydrobiopterin synthetic pathway and catecholamine biosynthesis.  相似文献   

8.
We have examined the interaction of hepatic phenylalanine hydroxylase with the phenylalanine analogs, tryptophan and the diastereomers of 3-phenylserine (beta-hydroxyphenylalanine). Both isomers of phenylserine are substrates for native phenylalanine hydroxylase at pH 6.8 and 25 degrees C, when activity is measured with the use of the dihydropteridine reductase assay coupled with NADH in the presence of the synthetic cofactor, 6-methyl-5,6,7,8-tetrahydropterin. However, while erythro-phenylserine exhibits simple Michaelis-Menten kinetics (Km = 1.2 mM, Vmax = 1.2 mumol/min X min) under these conditions, the threo isomer exhibits strong positive cooperativity (S0.5 = 4.8 mM Vmax = 1.4 mumol/min X mg, nH = 3). Tryptophan also exhibits cooperativity under these conditions (S0.5 = 5 mM, Vmax = 1 mumol/min X mg, nH = 3). The presence of 1 mM lysolecithin results in a hyperbolic response of phenylalanine hydroxylase to tryptophan (Km = 4 mM, Vmax = 1 mumol/min X mg) and threo-phenylserine (Km = 2 mM, Vmax = 1.4 mumol/min X mg). erythro-Phenylserine is a substrate for native phenylalanine hydroxylase in the presence of the natural cofactor, L-erythro-tetrahydrobiopterin (BH4) (Km = 2 mM, Vmax 0.05 mumol/min X mg, nH = 2). Preincubation of phenylalanine hydroxylase with erythro-phenylserine results in a 26-fold increase in activity upon subsequent assay with BH4 and erythro-phenylserine, and hyperbolic kinetic plots are observed. In contrast, both threo-phenylserine and tryptophan exhibit negligible activity in the presence of BH4 unless the enzyme has been activated. The product of the reaction of phenylalanine hydroxylase with either isomer of phenylserine was identified as the corresponding p-hydroxyphenylserine by reaction with sodium periodate and nitrosonaphthol. With erythro-phenylserine, the hydroxylation reaction is tightly coupled (i.e. 1 mol of hydroxyphenylserine is formed for every mole of tetrahydropterin cofactor consumed), while with threo-phenylserine and tryptophan the reaction is largely uncoupled (i.e. more cofactor consumed than product formed). Erythro-phenylserine is a good activator, when preincubated with phenylalanine hydroxylase (A0.5 = 0.2 mM), with a potency about one-third that of phenylalanine (A0.5 = 0.06 mM), while threo-phenylserine (A0.5 = 6 mM) and tryptophan (A0.5 approximately 10 mM) are very poor activators. Addition of 4 mM tryptophan or threo-phenylserine or 0.2 mM erythro-phenylserine to assay mixtures containing BH4 and phenylalanine results in a dramatic increase in the hydroxylation at low concentrations of phenylalanine.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

9.
This study has characterized two new enzymatic hydroxylase activities specific for 5 alpha-androstane-3 beta, 17 beta-diol (3 beta-diol) in the rat ventral prostate: 5 alpha-androstane-3 beta, 17 beta-diol 6 alpha-hydroxylase (6 alpha-hydroxylase) and 5 alpha-androstane-3 beta, 17 beta-diol 7 alpha-hydroxylase (7 alpha-hydroxylase). Both of these irreversible hydroxylase activities require NADPH and are localized in the microsomal fraction of the prostate. The apparent Km for 3 beta-diol is 2.5 microM for both the 6 alpha- and 7 alpha-hydroxylase activities. The apparent Km for NADPH is 7.6 microM for the 6 alpha-hydroxylase and 7.0 microM for the 7 alpha-hydroxylase. The pH optimum for both activities is 7.4. Several steroid inhibitors of these hydroxylase activities in vitro were identified including cholesterol, progesterone, and estradiol. Estradiol was found in vitro to be a noncompetitive inhibitor (Ki = 5 microM). Injection of estradiol into intact male rats, simultaneously receiving exogenous testosterone, also produced a significant lowering of the 6 alpha-plus 7 alpha-hydroxylase activities. Both the 6 alpha- and 7 alpha-hydroxylase were found to be androgen sensitive. Following castration there is a rapid decrease in both activities.  相似文献   

10.
In vivo measurements by positron emission tomography of the brain serotonin synthesis rates in the normal dog, in the dog with increased plasma tryptophan concentration, and in the dog under different arterial oxygen tensions are described. The method described here permits repeated measurements in the same brain for the first time. An increase in the plasma tryptophan concentration from 16.6 to 191.5 and then to 381 microM resulted in close to a linear increase in the brain serotonin synthesis rate. When PaO2 was raised from 76 +/- 2 to 106 +/- 1 mm Hg, the rate of serotonin synthesis in the dog brain increased from 39 +/- 8 to 54 +/- 10 pmol g-1 min-1. The estimates of the Michaelis-Menten constants, Kappm and Vmax, for the transport of tryptophan through the blood-brain barrier are 303 +/- 54 microM and 63 +/- 10 nmol g-1 min-1, respectively.  相似文献   

11.
Abstract— Circadian variations in the activity of tyrosine hydroxylase, tyrosine aminotransferase, and tryptophan hydroxylase were observed in the rat brain stem. Tyrosine hydroxylase exhibited a bimodal pattern with peaks occurring during both the light and dark phases of the circadian cycle. Tyrosine aminotransferase had one daily peak of activity occurring late in the light phase, whereas tryptophan hydroxylase activity was maximal late in the dark phase. Circadian fluctuations in tyrosine hydroxylase activity did not correlate well with circadian variations in the turnover rates of norepinephrine or dopamine nor with levels of these catecholamines. This supports the idea that although tyrosine hydroxylase is the rate-limiting enzyme in the synthesis of catecholamines, other factors must also be involved in the in vivo regulation of this process. Administration of α -methyl- p -tyrosine (AMT) methyl ester HC1 (100 mg/kg) had no effect on the activity of tryptophan hydroxylase, but effectively eliminated the peak of tyrosine hydroxylase activity that occurred during the light phase. AMT also lowered levels of tyrosine aminotransferase, but only at times near the daily light to dark transition. These chronotypic effects of AMT emphasize the importance of "time of day" as a factor that must be taken into account in evaluating the biochemical as well as the pharmacological and toxicological effects of drugs.  相似文献   

12.
In the presence of tyrosine, phenylalanine hydroxylase, which has been activated with lysolecithin, catalyzes the oxidation of tetrahydrobiopterin at a rate 10-20% that of the parallel reaction with phenylalanine. Unlike the reaction with phenylalanine, there is no net concomitant hydroxylation of tyrosine, although the amino acid is still a necessary component. Tyrosine appears to form an abortive complex with the activated enzyme, the pterin cofactor and molecular oxygen. The Km for tetrahydrobiopterin is identical for the reactions with phenylalanine and tyrosine, whereas the Km for tyrosine is approximately 3 1/2 times greater than the Km for phenylalanine. The tyrosine-dependent oxidation of tetrahydrobiopterin proceeds at both pH 6.8 and 8.2 and shows a similar dependence on the pH as that of the physiological reaction. Tetrahydrobiopterin can be replaced by the artificial cofactor, 6-methyltetrahydropterin, in the tyrosine-dependent oxidation at both pH 6.8 and 8.2. As in the parallel reaction with phenylalanine, both the Km for the cofactor and the Km for the aromatic amino acid increase with this substitution.  相似文献   

13.
Tryptophan uptake, hydroxylation, and decarboxylation in isolated synaptosomes were studied to assess how their properties may determine the rate of serotonin synthesis in the presynaptic nerve terminals of the brain. Simultaneous measurements of the rates of uptake, hydroxylation, and decarboxylation in the presence and absence of various inhibitors showed that tryptophan hydroxylase is rate-limiting for serotonin synthesis in this model system. There was significant direct decarboxylation of tryptophan to tryptamine. Measurement of tryptophan hydroxylase flux with varying internal concentrations of tryptophan allowed the determination of the Km of tryptophan hydroxylase in synaptosomes for tryptophan of 120 +/- 15 microM. Depolarisation of synaptosomes with veratridine caused both a reduction in the internal tryptophan concentration and an apparent activation of tryptophan hydroxylase. This activation did not occur in the absence of Ca2+ or in the presence of trifluoperazine. Synaptosomal serotonin synthesis and brain stem-soluble tryptophan hydroxylase were inhibited by low concentrations of noradrenaline or dopamine. Dibutyryl cyclic AMP, glucagon, insulin, and vasopressin were observed to have no effect on tryptophan uptake or hydroxylation in synaptosomes.  相似文献   

14.
Deacetoxycephalosporin C hydroxylase from cell-free extracts of Streptomyces clavuligerus was stabilized partially and purified to near homogeneity by three anion-exchange chromatographies, ammonium sulfate fractionation, and two gel filtrations. The hydroxylase was a monomer with a Mr of 35,000-38,000. alpha-Ketoglutarate, ferrous iron, and molecular oxygen were required for the enzyme activity. The hydroxylase was optimally active between pH 7.0 and 7.4 in a 3-(N-morpholino)propanesulfonic acid buffer and at 29 degrees C. It was stimulated by a reducing agent, particularly dithiothreitol or reduced glutathione, and ATP. The requirement for ferrous ion was specific, and at least one sulfhydryl group was apparently essential for the enzymatic hydroxylation. The Km values of the hydroxylase for deacetoxycephalosporin C and alpha-ketoglutarate were 59 and 10 microM, respectively, and the Ka for ferrous ion was 20 microM. In addition to its known hydroxylation of deacetoxycephalosporin C to deacetylcephalosporin C, the hydroxylase catalyzed effectively an analogous hydroxylation of 3-exomethylenecephalosporin C to deacetoxycephalosporin C. Surprisingly, the hydroxylase also mediated slightly a novel ring-expansion of penicillin N to deacetoxycephalosporin C. The substrate specificity of the hydroxylase is overlapping with but distinguishable from that of deacetoxycephalosporin C synthase, the enzyme which normally mediates the ring-expansion reaction (Dotzlaf, J. E., and Yeh, W. K. (1989) J. Biol. Chem. 264, 10219-10227). Furthermore, the hydroxylase exhibited an extensive sequence similarity to the synthase. Thus, the two enzymes catalyzing the consecutive reactions for cephamycin C biosynthesis in S. clavuligerus represent apparent products from a divergent evolution.  相似文献   

15.
We have studied the intrinsic fluorescence of the 12 tryptophan residues of electron-transfer flavoprotein:ubiquinone oxidoreductase (ETF:QO). The fluorescence emission spectrum (lambda ex 295 nm) showed that the fluorescence is due to the tryptophan residues and that the contribution of the 22 tyrosine residues is minor. The emission maximum (lambda m 334 nm) and the bandwidth (delta lambda 1/2 56 nm) suggest that the tryptophans lie in hydrophobic environments in the oxidized protein. Further, these tryptophans are inaccessible to a range of ionic and nonionic collisional quenching agents, indicating that they are buried in the protein. Enzymatic or chemical reduction of ETF:QO results in a 5% increase in fluorescence with no change of lambda m or delta lambda 1/2. This change is reversible upon reoxidation and is likely to reflect a conformational change in the protein. The ubiquinone analogue Q0(CH2)10Br, a pseudosubstrate of ETF:QO (Km = 2.6 microM; kcat = 210 s-1), specifically quenches the fluorescence of one tryptophan residue (Kd = 1.6-3.2 microM) in equilibrium fluorescence titrations. The ubiquinone homologue UQ-2 (Km = 2 microM; kcat = 162 s-1) and the analogue Q0(CH2)10OH (Km = 2 microM; kcat = 132 s-1) do not quench tryptophan fluorescence; thus the brominated analogue acts as a static heavy atom quencher. We also describe a rapid purification for ETF:QO based on extraction of liver submitochondrial particles with Triton X-100 and three chromatographic steps, which results in yields 3 times higher than previously published methods.  相似文献   

16.
Tyrosine hydroxylase in bovine adrenal medulla was activated up to fourfold by incubation with low concentrations (15 micrograms/ml) of ribonucleic acids. At higher RNA concentrations, enzyme activity was inhibited. This interaction with RNA was exploited with the use of poly(A)-Sepharose and DNA-cellulose to effect a rapid purification of stable tyrosine hydroxylase from rat brain and bovine adrenal medulla in high yield (up to 58%). With the purified rat brain enzyme, RNA acted as an uncompetitive inhibitor, a concentration of 15 micrograms/ml lowering the Vmax of tyrosine hydroxylase from 1050 to 569 nmol min-1 mg-1 and lowering the Km for tyrosine from 6.1 to 3.6 microM. With the natural cofactor, tetrahydrobiopterin (BH4), two Km values were obtained, indicating the presence of two forms of the enzyme. Both Km values were decreased only slightly by RNA. The purified brain and adrenal enzymes both contained about 0.07 mol of phosphate/63,000-Da subunit; in both cases, cyclic AMP-dependent protein kinase catalyzed the incorporation of an additional 0.8 mol of phosphate/subunit. The purified enzyme also contains ribonucleic acid, which comprises about 10% of the total mass and appears to be important for full activity.  相似文献   

17.
Tyrosine hydroxylase activity has been established in blood plasma leucocytes of rat, cat and man. Tyrosine precursors and some nuclear erythroid cells. GFU-GM did hydroxylase activity in leucocytes shows the Km for tyrosine inhibited by high concentrations of L6 tyrosine (substrate inhibition), alpha-methyl-para-tyrosine dopamine. The kinetic properties of leucocyte tyrosine hydroxylase are qualitatively similar to the properties of brain tyrosine hydroxylase.  相似文献   

18.
G J Lees 《Life sciences》1977,20(10):1749-1762
Using low concentrations of substrates and cofactors, a comparison was made of the relative rates by which aminotransferases catalysed transaminations between aromatic amino acids and aromatic or aliphatic keto acids. Tryptophan aminotransferase in homogenates of rat midbrain and liver transaminated phenylpyruvate at a rate 70 to 150-fold greater than the rate with α-ketoglutarate at low concentrations of substrates. Phenylalanine aminotransferase in liver and midbrain also was more active with aromatic keto acids than with aliphatic keto acids. However, tyrosine aminotransferase in dialysed homogenates of midbrain transaminated α-ketoglutarate and phenylpyruvate at approximately equal rates. Fresh homogenates of midbrain contained an inhibitor which markedly decreased tyrosine aminotransferase activity with α-ketoglutarate but not with phenylpyruvate. Tyrosine aminotransferase in homogenates of rat liver transaminated α-ketoglutarate and phenylpyruvate at equal rates below 10 μM keto acid, but above 10 μM, transamination of α-ketoglutarate was favoured. With homogenates of liver, transamination of α-ketoglutarate, but not phenylpyruvate, by tyrosine was increased 650% by exogenous pyridoxal phosphate. Since tryptophan aminotransferase in the brain may compete with tryptophan hydroxylase for available tryptophan, a comparison was made of the relative activities of tryptophan hydroxylase and tryptophan aminotransferase. At concentrations above 7.5 μM phenylpyruvate, transamination was 8 to 17-fold greater than the rate of hydroxylation of 50 μM tryptophan.  相似文献   

19.
NADPH-cytochrome P-450 (cytochrome c) reductase (EC 1.6.2.4) was solubilized by detergent from microsomal fraction of wounded Jerusalem-artichoke (Helianthus tuberosus L.) tubers and purified to electrophoretic homogeneity. The purification was achieved by two anion-exchange columns and by affinity chromatography on 2',5'-bisphosphoadenosine-Sepharose 4B. An Mr value of 82,000 was obtained by SDS/polyacrylamide-gel electrophoresis. The purified enzyme exhibited typical flavoprotein redox spectra and contained equimolar quantities of FAD and FMN. The purified enzyme followed Michaelis-Menten kinetics with Km values of 20 microM for NADPH and 6.3 microM for cytochrome c. In contrast, with NADH as substrate this enzyme exhibited biphasic kinetics with Km values ranging from 46 microM to 54 mM. Substrate saturation curves as a function of NADPH at fixed concentration of cytochrome c are compatible with a sequential type of substrate-addition mechanism. The enzyme was able to reconstitute cinnamate 4-hydroxylase activity when associated with partially purified tuber cytochrome P-450 and dilauroyl phosphatidylcholine in the presence of NADPH. Rabbit antibodies directed against plant NADPH-cytochrome c reductase affected only weakly NADH-sustained reduction of cytochrome c, but inhibited strongly NADPH-cytochrome c reductase and NADPH- or NADH-dependent cinnamate hydroxylase activities from Jerusalem-artichoke microsomal fraction.  相似文献   

20.
The structure of the cofactor binding domain of tyrosine hydroxylase (TH) was examined at physiological pH by determining kinetic parameters of (R)-tetrahydrobiopterin [(R)-BH4] and a series of tetrahydropterin (PH4) derivatives (6-R1-6-R2-PH4: R1 = H and R2 = methyl, hydroxymethyl, ethyl, methoxymethyl, phenyl, and cyclohexyl; R1 = methyl and R2 = methyl, ethyl, propyl, phenyl, and benzyl). A minimally purified TH preparation that was not specifically phosphorylated (designated as "unphosphorylated") was compared with enzyme phosphorylated with cAMP-dependent protein kinase. The Km for tyrosine with most tetrahydropterin analogues ranged between 20 and 60 microM with little decrease upon phosphorylation. Two exceptions were an unusually low Km of 7 microM with 6-ethyl-PH4 and a high Km of 120 microM with 6-phenyl-6-methyl-PH4, both with phosphorylated TH. Tyrosine substrate inhibition was elicited only with (R)-BH4 and 6-hydroxymethyl-PH4. With unphosphorylated TH (with the exception of 6-benzyl-6-methyl-PH4, Km = 4 mM) an inverse correlation between cofactor Km and side-chain hydrophobicity was observed ranging from a high with (R)-BH4 (5 mM) to a low with 6-cyclohexyl-PH4 (0.3 mM). An 8-fold span of Vmax was seen overall. Phosphorylation caused a 0.6-4-fold increase in Vmax and a 35-2000-fold decrease in Km for cofactor, ranging from a high of 60 microM with 6-methyl-PH4 to a low of 0.6 microM with 6-cyclohexyl-PH4. A correlation of the size of the hydrocarbon component of the side chain with affinity is strongly evident with phosphorylated TH, but in contrast to unphosphorylated enzyme, the hydroxyl groups in hydroxymethyl-PH4 (20 microM) and (R)-BH4 (3 microM) decrease Km in comparison to that of 6-methyl-PH4. Although 6,6-disubstituted analogues were found with affinities near that of (R)-BH4 (e.g., 6-propyl-6-methyl-PH4, 4 microM), they were frequently more loosely associated with phosphorylated TH than their monosubstituted counterparts (6-phenyl-PH4, 0.8 microM; cf. 6-phenyl-6-methyl-PH4, 8 microM). A model of the cofactor side-chain binding domain is proposed in which a limited region of nonpolar protein residue(s) capable of van der Waals contact with the hydrocarbon backbone of the (R)-BH4 dihydroxypropyl group is opposite to a recognition site for hydroxyl(s). Although interaction with either the hydrophilic or hydrophobic regions of unphosphorylated tyrosine hydroxylase is possible, phosphorylation by cAMP-dependent protein kinase appears to optimize the simultaneous operation of both forces.  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