首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
A liquid chromatographic-fluorometric assay has been developed to study the role of dihydrofolate reductase in adult rat brain since low levels of the enzyme preclude measurement by current spectrophotometric procedures. This method involves in vitro incubation of desalted, cell-free brain extracts with 7,8-dihydrobiopterin, NADPH, and an NADPH-regenerating system. The tetrahydrobiopterin formed is quantitatively converted to pterin using alkaline iodine oxidation, and the pterin formed is separated by liquid chromatography and detected fluorometrically. The method is linear from 100 fmol to greater than or equal to 1 nmol of product, and the sensitivity is at least 100 times greater than that of existing spectrophotometric assays. Enzyme activity of desalted brain extracts is linear with both time (to 100 min) and protein (from 50 to 620 micrograms). The enzyme shows an absolute requirement for NADPH, does not use NADH, and is completely inhibited by 10 nM methotrexate. The Km of the enzyme for NADPH was found to be 7.5 microM, while the Km for 7,8-dihydrobiopterin was 88 microM. Since brain dihydrobiopterin reductase has the same properties as dihydrofolate reductase, this fluorometric procedure can serve as a sensitive assay for dihydrofolate reductase.  相似文献   

2.
Abstract: Culture of rat pineal glands in methotrexate (0.5, 5, or 10 μM) for 6 or 24 h did not alter pineal tetrahydrobiopterin (85–90% of total biopterin in cultured glands), except for a decrease of 30% after 24 h culture in 10 μM methotrexate. However, pineal dihydrobiopterin and/or biopterin (10–15% of total biopterin) was increased by methotrexate up to 2.5-fold. Biopterin detected in the culture medium following pineal culture was also increased to a similar extent after methotrexate treatment and appeared to represent leakage of pineal dihydrobiopterin and/or biopterin. Culture of glands in 5 μM methotrexate did not alter the conversion of [U-14C]-guanosine to [14C]biopterin, suggesting that pineal tetrahydrobiopterin synthesis was not altered by methotrexate. Complete inhibition of dihydrofolate reductase activity measured in pineal homogenates was obtained following culture of glands in all concentrations of methotrexate studied. Therefore, dihydrofolate reductase and dihydrobiopterin do not appear to be involved in a major biosynthetic pathway for pineal tetrahydrobiopterin from GTP, although they may have a minor role in tetrahydrobiopterin synthesis.  相似文献   

3.
A liquid chromatographic-fluorometric assay has been developed to study the role of dihydrofolate reductase in adult rat brain since low levels of the enzyme preclude measurement by current spectrophotometric procedures. This method involves in vitro incubation of desalted, cell-free brain extracts with 7,8-dihydrobiopterin, NADPH, and an NADPH-regenerating system. The tetrahydrobiopterin formed is quantitatively converted to pterin using alkaline iodine oxidation, and the pterin formed is separated by liquid chromatography and detected fluorometrically. The method is linear from 100 fmol to ≥ 1 nmol of product, and the sensitivity is at least 100 times greater than that of existing spectrophotometric assays. Enzyme activity of desalted brain extracts is linear with both time (to 100 min) and protein (from 50 to 620 μg). The enzyme shows an absolute requirement for NADPH, does not use NADH, and is completely inhibited by 10 n methotrexate. The Km of the enzyme for NADPH was found to be 7.5 μ , while the Km for 7,8-dihydrobiopterin was 88 μ . Since brain dihydrobiopterin reductase has the same properties as dihydrofolate reductase, this fluorometric procedure can serve as a sensitive assay for dihydrofolate reductase.  相似文献   

4.
Conditions for the quantitative oxidation and destruction of tetrahydrobiopterin and quinoid dihydrobiopterin and the separation of biopterin from its reduced forms by ECTEOLA-Sephadex column chromatography are described. A procedure for the quantitation of tetrahydrobiopterin plus quinoid dihydrobiopterin, 7,8-dihydrobiopterin, and biopterin using a Crithidia bioassay is presented. Using these procedures it was found that tetrahydrobiopterin plus quinoid dihydrobiopterin are the prevalent forms in liver and blood of mice and that biopterin was the predominant form in the tails of tadpoles. In human urine, approximately half of the biopterin was found as tetrahydrobiopterin plus quinoid dihydrobiopterin and the other half was 7,8-dihydrobiopterin. The presence of tetrahydrobiopterin and quinoid dihydrobiopterin was confirmed by a coenzyme assay for the hydroxylation of phenylalanine.  相似文献   

5.
The enzyme, D-erythrodihydroneopterin triphosphate synthetase from rat brain was observed to have a significantly lower specific activity than that from liver due to their degree of dephosphorylation during preparation. The brain enzyme could be phosphorylated in vitro in presence of [32P]-ATP and protein kinase, resulting in an increased specific activity. Isolation of brain enzyme in presence of 0.8 M NaF allowed recovery of the enzyme phosphorylated at residue 67 (serine) as determined by a new assay for phosphate. This enzyme is present in synaptosomes and its state of phosphorylation may regulate the rate at which dihydrobiopterin, the precursor of the hydroxylase cofactor (tetrahydrobiopterin, BH4), is synthesized by synaptosomes.  相似文献   

6.
A cDNA clone (SSC801) putatively encoding sepiapterin reductase (SR) was obtained from the expressed sequence tag clones of Dictyostelium discoideum. The cDNA sequence of 878 nucleotides constituted an ORF of 265 amino acid residues but was missing a few N-terminal residues. The deduced amino acid sequence showed 29.8% identity with mouse SR sequence and a molecular mass of 29,969 Da. The coding sequence was cloned in E. coli expression vector and overexpressed. The purified His-tag recombinant enzyme was confirmed to have the genuine activity of SR to produce tetrahydrobiopterin from 6-pyruvoyltetrahydropterin in a coupled assay with 6-pyruvoyltetrahydropterin synthase as well as dihydrobiopterin from sepiapterin. However, dictyopterin was not observed in our assay condition. The enzyme was also inhibited by N-acetylserotonin and to a lesser extent by melatonin. Km values for NADPH and sepiapterin were 51.8+/-2.7 microM and 40+/-2 microM, respectively. Vmax was determined as 0.14 micromol/min/mg of protein.  相似文献   

7.
To identify the new targets for hypertension, we analyzed the protein expression profiles of aortic smooth muscle in spontaneously hypertensive rats (SHR) of various ages during the development of hypertension, as well as in age‐matched normotensive Wistar–Kyoto (WKY) rats, using a proteomic analysis. The expressions of seven proteins were altered in SHR compared with WKY rats. Of these proteins, NADH dehydrogenase 1α, GSTω1, peroxi‐redoxin I and transgelin were upregulated in SHR compared with WKY rats. On the other hand, the expression of HSP27 and Ran protein decreased in SHR. The diminution of dihydrobiopterin reductase, an enzyme located in the regeneration pathways of tetrahydrobiopterin (BH4), was also prominent in SHR. The results from a PCR analysis revealed that the expression of BH4 biosynthesis enzymes – GTP cyclohydrolase‐1 and sepiapterin reductase – decreased and increased, respectively, in SHR compared with WKY rats. The level of BH4 was less in aortic strips from SHR than from WKY rats. Moreover, treatment with BH4 inhibited aortic smooth muscle contraction induced by serotonin. These results suggest that the deficiency in BH4 regeneration produced by diminished dihydrobiopterin reductase expression is involved in vascular disorders in hypertensive rats.  相似文献   

8.
Classic tetrahydrobiopterin (BH(4)) deficiencies are characterized by hyperphenylalaninemia and deficiency of monoamine neurotransmitters. In this article, we report two patients with progressive psychomotor retardation, dystonia, severe dopamine and serotonin deficiencies (low levels of 5-hydroxyindoleacetic and homovanillic acids), and abnormal pterin pattern (high levels of biopterin and dihydrobiopterin) in cerebrospinal fluid. Furthermore, they presented with normal urinary pterins and without hyperphenylalaninemia. Investigation of skin fibroblasts revealed inactive sepiapterin reductase (SR), the enzyme catalyzing the final two-step reaction in the biosynthesis of BH(4). Mutations in the SPR gene were detected in both patients and their family members. One patient was homozygous for a TC-->CT dinucleotide exchange, predicting a truncated SR (Q119X). The other patient was a compound heterozygote for a genomic 5-bp deletion (1397-1401delAGAAC) resulting in abolished SPR-gene expression and an A-->G transition leading to an R150G amino acid substitution and to inactive SR as confirmed by recombinant expression. The absence of hyperphenylalaninemia and the presence of normal urinary pterin metabolites and of normal SR-like activity in red blood cells may be explained by alternative pathways for the final two-step reaction of BH(4) biosynthesis in peripheral and neuronal tissues. We propose that, for the biosynthesis of BH(4) in peripheral tissues, SR activity may be substituted by aldose reductase (AR), carbonyl reductase (CR), and dihydrofolate reductase, whereas, in the brain, only AR and CR are fully present. Thus, autosomal recessive SR deficiency leads to BH(4) and to neurotransmitter deficiencies without hyperphenylalaninemia and may not be detected by neonatal screening for phenylketonuria.  相似文献   

9.
Enzymology of the phenylalanine-hydroxylating system   总被引:1,自引:0,他引:1  
S Kaufman 《Enzyme》1987,38(1-4):286-295
The phenylalanine-hydroxylating system consists of 3 essential components, phenylalanine hydroxylase (PAH), dihydropteridine reductase (DHPR) and the coenzyme, tetrahydrobiopterin (BH4). DHPR and BH4 are also essential components of the trosine- and tryptophan-hydroxylating systems. During the hydroxylation reaction, BH4 is converted to the quinonoid dihydrobiopterin. The reduction of this latter compound back to BH4 is catalyzed by the reductase in the presence of NADH. In addition to the classic form of phenylketonuria, which is caused by a lack of PAH, a form is caused by a lack of DHPR and another by a deficiency of BH4 caused by the lack of an enzyme involved in its de novo biosynthesis. Besides hyperphenylalaninemia, these variant forms are characterized by neurological deterioration.  相似文献   

10.
The reduction of 7,8-dihydrobiopterin to 5,6,7,8-tetrahydrobiopterin by rat liver tetrahydrofolate dehydrogenase (5,6,7,8-tetrahydrofolate-NADP+ oxidoreductase, EC 1.5.1.3) is competitively inhibited by trimethoprim lactate (apparent Ki 0.285 muM). An apparent Michaelis constant of 43 muM for dihydrobiopterin was obtained, which is 430 times higher than the reported Km for dihydrofolate with this enzyme. The reduction of dihydrobiopterin is thus more susceptible to inhibition by trimethoprim lactate than is the reduction of dihydrofolate. However, intraperitoneal administration of trimethoprim had no significant effect on the hepatic supply of tetrahydrobiopterin in rats.  相似文献   

11.
The biosynthesis of tetrahydrobiopterin from either dihydroneopterin triphosphate, sepiapterin, dihydrosepiapterin or dihydrobiopterin was investigated using extracts from human liver, dihydrofolate reductase and purified sepiapterin reductase from human liver and rat erythrocytes. The incorporation of hydrogen in tetrahydrobiopterin was studied in either 2H2O or in H2O using unlabeled NAD(P)H or (R)-(4-2H)NAD(P)H or (S)-(4-2H)NAD(P)H. Dihydrofolate reductase catalyzed the transfer of the pro-R hydrogen of NAD(P)H during the reduction of 7,8-dihydrobiopterin to tetrahydrobiopterin. Sepiapterin reductase catalyzed the transfer of the pro-S hydrogen of NADPH during the reduction of sepiapterin to 7,8-dihydrobiopterin. In the presence of partially purified human liver extracts one hydrogen from the solvent is introduced at position C(6) and the 4-pro-S hydrogen from NADPH is incorporated at each of the C(1') and C(2') position of BH4. Label from the solvent is also introduced into position C(3'). These results suggest that dihydrofolate reductase is not involved in the biosynthesis of tetrahydrobiopterin from dihydroneopterin triphosphate. They are consistent with the assumption of the occurrence of a 6-pyruvoyl-tetrahydropterin intermediate, which is proposed to be formed upon triphosphate elimination from dihyroneopterin triphosphate, and via an intramolecular redox reaction. Our results suggest that the reduction of 6-pyruvoyl-tetrahydropterin might be catalyzed by sepiapterin reductase.  相似文献   

12.
An enzyme with 6-pyruvoyl tetrahydropterin (6PPH4) (2'-oxo)reductase activity was purified to near homogeneity from whole rat brains by a rapid method involving affinity chromatography on Cibacron blue F3Ga-agarose followed by high performance ion exchange chromatography and high performance gel filtration. The enzyme has a single subunit of Mr 37,000 and has a similar amino acid composition to previously described aldoketo reductases. The reductase activity is absolutely dependent on NADPH, will only catalyze the reduction of the C-2'-oxo group of 6PPH4, and is inactive towards the C-1'-oxo group. However, the enzyme also shows high activity towards nonspecific substrates, such as 4-nitrobenzaldehyde, phenanthrenequinone, and menadione. The role of this 6PPH4 reductase in the formation of tetrahydrobiopterin (BH4) was investigated. Measurements were made of the rate of conversion of 6PPH4, generated from dihydroneopterin triphosphate with purified 6PPH4 synthase, to BH4 in the presence of mixtures of pure sepiapterin reductase and the 6PPH4 (2'-oxo)reductase purified from rat brains. The results suggest that when sepiapterin reductase activity is limiting, a large proportion of BH4 synthesis proceeds through the 6-lactoyl intermediate. However, when sepiapterin reductase is not limiting, most of the BH4 is probably formed via reduction of the other mono-reduced intermediate which is produced from 6PPH4 by sepiapterin reductase alone.  相似文献   

13.
Abstract: Tryptophan hydroxylase, the initial and rate-limiting enzyme in the biosynthesis of the neurotransmitter serotonin, is inactivated by the nitric oxide generators sodium nitroprusside, diethylamine/nitric oxide complex, and S -nitroso- N -acetylpenicillamine. Physiological concentrations of tetrahydrobiopterin, the natural and endogenous cofactor for the hydroxylase, significantly enhance the inactivation of the enzyme caused by each of these nitric oxide generators. The substrate tryptophan does not have this effect. The chemically reduced (tetrahydro-) form of the pterin is required for the enhancement, because neither biopterin nor dihydrobiopterin is effective. The 6 S -isomer of tetrahydrobiopterin, which has little cofactor efficacy for tryptophan hydroxylase, does not enhance enzyme inactivation as does the natural 6 R -isomer. A number of synthetic, reduced pterins share with tetrahydrobiopterin the ability to enhance nitric oxide-induced inactivation of tryptophan hydroxylase. The tetrahydrobiopterin effect is not prevented by agents known to scavenge hydrogen peroxide, superoxide radicals, peroxynitrite anions, hydroxyl radicals, or singlet oxygen. On the other hand, cysteine partially protects the enzyme from both the nitric oxide-induced inactivation and the combined pterin/nitric oxide-induced inactivation. These results suggest that the tetrahydrobiopterin cofactor enhances the nitric oxide-induced inactivation of tryptophan hydroxylase via a mechanism that involves attack on free protein sulfhydryls. Potential in vivo correlates of a tetrahydrobiopterin participation in the inactivation of tryptophan hydroxylase can be drawn to the neurotoxic amphetamines.  相似文献   

14.
Leishmania parasites are pteridine auxotrophs that use an NADPH-dependent pteridine reductase 1 (PTR1) and NADH-dependent quinonoid dihydropteridine reductase (QDPR) to salvage and maintain intracellular pools of tetrahydrobiopterin (H(4)B). However, the African trypanosome lacks a credible candidate QDPR in its genome despite maintaining apparent QDPR activity. Here we provide evidence that the NADH-dependent activity previously reported by others is an assay artifact. Using an HPLC-based enzyme assay, we demonstrate that there is an NADPH-dependent QDPR activity associated with both TbPTR1 and LmPTR1. The kinetic properties of recombinant PTR1s are reported at physiological pH and ionic strength and compared with LmQDPR. Specificity constants (k(cat)/K(m)) for LmPTR1 are similar with dihydrobiopterin (H(2)B) and quinonoid dihydrobiopterin (qH(2)B) as substrates and about 20-fold lower than LmQDPR with qH(2)B. In contrast, TbPTR1 shows a 10-fold higher k(cat)/K(m) for H(2)B over qH(2)B. Analysis of Trypanosoma brucei isolated from infected rats revealed that H(4)B (430 nM, 98% of total biopterin) was the predominant intracellular pterin, consistent with a dual role in the salvage and regeneration of H(4)B. Gene knock-out experiments confirmed this: PTR1-nulls could only be obtained from lines overexpressing LmQDPR with H(4)B as a medium supplement. These cells grew normally with H(4)B, which spontaneously oxidizes to qH(2)B, but were unable to survive in the absence of pterin or with either biopterin or H(2)B in the medium. These findings establish that PTR1 has an essential and dual role in pterin metabolism in African trypanosomes and underline its potential as a drug target.  相似文献   

15.
The NADPH-dependent reduction of the two carbonyl groups in the side chain of the first tetrahydropterin intermediate on the tetrahydrobiopterin biosynthetic pathway, 6-pyruvoyl tetrahydropterin, proceeds in a sequential manner whose order has not yet been resolved. Sepiapterin reductase can catalyze the reduction of both carbonyl groups starting with the 1'-oxo. 6-Pyruvoyl tetrahydropterin (2'-oxo) reductase, which has now been shown to be a member of the aldose reductase family, catalyzes the formation of only the 2'-hydroxy-1'-oxo intermediate which still requires sepiapterin reductase for final conversion to tetrahydrobiopterin. Inhibiting antibodies to the 2'-oxo reductase have been prepared and utilized to explore the distribution of this reductase in rat brain. The antiserum also maximally inhibited in vitro tetrahydrobiopterin synthesis in crude rat brain extracts by 60%, indicating that the majority of tetrahydrobiopterin biosynthesis in vivo may proceed via the 2'-hydroxy-1'-oxo intermediate. However, analogous experiments with rat liver extracts demonstrate that inhibition of the 2'-oxo reductase activity does not inhibit the conversion of 6-pyruvoyl tetrahydropterin to tetrahydrobiopterin, suggesting that tetrahydrobiopterin biosynthesis may proceed via different pathways in rat brain and liver.  相似文献   

16.
Selective modification of the tetrahydrobiopterin levels in cultured chromaffin cells were followed by changes in the rate of tyrosine hydroxylation. Addition of sepiapterin, an intermediate on the salvage pathway for tetrahydrobiopterin synthesis, rapidly increased intracellular levels of tetrahydrobiopterin and elevated the rate of tyrosine hydroxylation in the intact cell. Tyrosine hydroxylation was also enhanced when tetrahydrobiopterin was directly added to the incubation medium of intact cells. When the cultured chromaffin cells were treated for 72 h with N-acetylserotonin, an inhibitor of sepiapterin reductase, tetrahydrobiopterin content and the rate of tyrosine hydroxylation were decreased. Addition of sepiapterin or N-acetylserotonin had no consistent effect on total extractable tyrosine hydroxylase activity or on catecholamine content in the cultured chromaffin cells. Three-day treatment of chromaffin cell cultures with compounds that increase levels of cyclic AMP (forskolin, cholera toxin, theophylline, dibutyryl- and 8-bromo cyclic AMP) increased total extractable tyrosine hydroxylase activity and GTP-cyclohydrolase, the rate-limiting enzyme in the biosynthesis of tetrahydrobiopterin. Tetrahydrobiopterin levels and intact cell tyrosine hydroxylation were markedly increased after 8-bromo cyclic AMP. The increase in GTP-cyclohydrolase and tetrahydrobiopterin induced by 8-bromo cyclic AMP was blocked by the protein synthesis inhibitor cycloheximide. Agents that deplete cellular catecholamines (reserpine, tetrabenazine, and brocresine) increased both total tyrosine hydroxylase and GTP-cyclohydrolase activities, although treating the cultures with reserpine or tetrabenazine resulted in no change in cellular levels of cyclic AMP. Brocresine and tetrabenazine increased tetrahydrobiopterin levels, but the addition of reserpine to the cultures decreased catecholamine and tetrahydrobiopterin content and resulted in a decreased rate of intact cell tyrosine hydroxylation in spite of the increased activity of the total extractable enzyme. These data indicate that in cultured chromaffin cells GTP-cyclohydrolase activity like tyrosine hydroxylase activity is regulated by both cyclic AMP-dependent and cyclic AMP-independent mechanisms and that the intracellular level of tetrahydrobiopterin is one of the many factors that control the rate of tyrosine hydroxylation.  相似文献   

17.
Sepiapterin reductase from rat erythrocyte hemolysate was purified 2000-fold to apparent homogeneity with 30% yield. The specific activity of the purified enzyme was 18 units/mg protein, and its molecular weight was 55 000. The enzyme consists of two identical subunits, each of which has a molecular weight of 27 500. The enzyme showed a single peak by isoelectric focusing with a pI of 4.9 and partial specific volume of 0.73 cm3/g. The amino acid composition was determined. pH optimum of the enzyme was 5.5. The equilibrium constant of 2.2·109 of the enzyme showed that the equilibrium lies much in favor of dihydrobiopterin formation from sepiapterin in rat erythrocytes. From steady-state kinetic measurements, ordered bi-bi mechanism was proposed to the reaction of sepiapterin reductase in which NADPH binds to free enzyme and sepiapterin binds next. NADP+ is released after the release of dihydrobiopterin. The Km values for sepiapterin and NADPH were 15.4 μM and 1.7 μM, respectively, and the Vmax value was 21.7 μmol/min per mg.  相似文献   

18.
A method was developed to study the unsupplemented phenylalanine hydroxylase system in rat liver slices. All of the components of the system--tetrahydrobiopterin, dihydropteridine reductase, and the hydroxylase itself--are present under conditions which should be representative of the actual physiological state of the animal. The properties of the system in liver slices have been compared to those of the purified enzyme in vitro. The three pterins, tetrahydrobiopterin, 6,7-dimethyltetrahydropterin, and 6-methyltetrahydropterin, all stimulate the hydroxylation of phenylalanine when added to the liver slice medium in the presence of a chemical reducing agent. The relative velocities found at 1 mM phenylalanine and saturating pterin concentrations are: tetrahydrobiopterin, 1; 6,7-dimethyltetrahydropterin, 2.5; 6-methyltetrahydropterin, 13. This ratio of activities is similar to that found for the purified, native phenylalanine hydroxylase and indicates that the enzyme in vivo is predominantly in the native form. Rats pretreated with 6-methyltetrahydropterin showed enhanced phenylalanine hydroxylase activity in liver slices demonstrating for the first time that an exogenous tetrahydropterin can interact with the phenylalanine hydroxylase system in vivo. This finding opens up the possibility of treating phenylketonurics who still possess some residual phenylalanine hydroxylase activity with a tetrahydropterin like 6-methyltetrahydropterin which can give a large increase in rate over that seen with the natural cofactor, tetrahydrobiopterin.  相似文献   

19.
Tetrahydrobiopterin plays an important role in the biosynthesis of certain neurotransmitters. Using DEAE-Sepharose FF column chromatography, we separated the enzyme that synthesizes tetrahydrobiopterin from 6-pyruvoyl-tetrahydropterin [which is different from sepiapterin reductase (EC 1.1.1.153)] in the lemon mutant of the silkworm Bombyx mori into two fractions, which were named carbonyl reductase I (CR I) and carbonyl reductase II (CR II). The CR I enzyme converted 6-pyruvoyl-tetrahydropterin to 6-lactoyl-tetrahydropterin, while CR II converted 6-pyruvoyl-tetrahydropterin to 1'-hydroxy-2'-oxopropyl-tetrahydropterin, both reactions occurring only in the presence of NADPH. Neither of the two carbonyl reductases alone was able to catalyze the conversion of 6-pyruvoyl-tetrahydropterin to tetrahydrobiopterin in the presence of NADPH. However, when CR I was mixed with CR II in the reaction mixture, 6-pyruvoyl-tetrahydropterin was reduced to tetrahydrobiopterin in the presence of NADPH. Moreover, CR I catalyzed the formation of tetrahydrobiopterin from 1'-hydroxy-2'-oxopropyl-tetrahydropterin, while CR II converted 6-lactoyl-tetrahydropterin to tetrahydrobiopterin, both reactions occurring only in the presence of NADPH. Our results suggest that there are two potential routes for formation of tetrahydrobiopterin from 6-pyruvoyl-tetrahydropterin in the lemon mutant silkworm. In the first route, 1'-hydroxy-2'-oxopropyl-tetrahydropterin is formed from 6-pyruvoyl-tetrahydropterin by CR II and then reduced to tetrahydrobiopterin by CR I, both reactions occurring only in the presence of NADPH. In the other route, 6-pyruvoyl-tetrahydropterin is reduced to 6-lactoyl-tetrahydropterin by CR I and then converted to tetrahydrobiopterin by CR II, both reactions occurring only in the presence of NADPH.  相似文献   

20.
The uncoupled portion of the partially uncoupled oxidation of tetrahydropterins by phenylalanine hydroxylase can be described by the same model as we have recently derived for the fully uncoupled reaction (Davis, M.D. and Kaufman, S. (1989) J. Biol. Chem.264, 8585–8596). Although essentially no hydrogen peroxide is formed during the fully coupled oxidation of tetrahydrobiopterin or 6-methyltetrahydropterin by phenylalanine hydroxylase when phenylalanine is the amino acid substrate, significant amounts of hydrogen peroxide are formed during the partially uncoupled oxidation of 6-methyltetrahydropterin whenpara-fluorophenylalanine orpara-chlorophenylalanine are used in place of phenylalanine. Similarly, during the partially uncoupled oxidation of the unsubstituted pterin, tetrahydropterin, even in the presence of phenylalanine, hydrogen peroxide formation is detected. The 4a-carbinolamine tetrahydropterin intermediate has been observed during the fully uncoupled tyrosine-dependent oxidations of tetrahydropterin and 6-methyltetrahydropterin by lysolecithin-activated phenylalanine hydroxylase, suggesting that this species is also a common intermediate for uncoupled oxidations by this enzyme.Abbreviations BH4 6-[dihydroxypropyl-(L-erythro)-5,6,7,8-tetrahydropterin (tetrahydrobiopterin) - 6MPH4 6-methyl-5,6,7,8-tetrahydropterin - PH4 5,6,7,8-tetrahydropterin - BH3OH 4a-hydroxytetrahydropterin (4a-carbinolamine) - qBH2 quinonoid dihydrobiopterin - q6MPH2 quinonoid dihydro-6-methylpterin - qPH2 quinoid dihydropterin - PAH phenylalanine hydroxylase - DHPR dihydropteridine reductase - PHS phenylalanine hydroxylase stimulating enzyme which is 4a-carbinolamine dehydratase - SOD superoxide dismutase - HPLC high performance liquid chromatography - R.T. retention time Special issue dedicated to Dr. Santiago Grisolia.  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