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.
The substrates of dihydropteridine reductase (EC 1.6.99.7), quinonoid 7,8-dihydro(6 H)pterins, are unstable and decompose in various ways. In attempting to prepare a more stable substrate, 6,6,8-trimethyl-5,6,7,8-tetrahydro(3 H)pterin was synthesised and the quinonoid 6,6,8-trimethyl-7,8-dihydro(6 H)pterin derived from it is extremely stable with a half-life in 0.1 M Tris/HCl (pH 7.6, 25 degrees C) of 33 h. Quinonoid 6,6,8-trimethyl-7,8-dihydro(6 H)pterin is not a substrate for dihydropteridine reductase but it is reduced non-enzymically by NADH at a significant rate and it is a weak inhibitor of the enzyme: I50 200 microM, pH 7.6, 25 degrees C when using quinonoid 6-methyl-7,8-dihydro(6 H)pterin as substrate. 6,6,8-Trimethyl-5,6,7,8-tetrahydropterin is a cofactor for phenylalanine hydroxylase (EC 1.14.16.1) with an apparent Km of 0.33 mM, but no cofactor activity could be detected with tyrosine hydroxylase (EC 1.14.16.2). Its phenylalanine hydroxylase activity, together with the enhanced stability of quinonoid 6,6,8-trimethyl-7,8-dihydro(6 H)pterin, suggest that it may have potential for the treatment of variant forms of phenylketonuria.  相似文献   

2.
The Km and kcat. values for [6,6,7,7-2H]7,8(6H)-dihydropterin and 2,6-diamino-5-iminopyrimidin-4-one were determined for dihydropteridine reductase (EC 1.6.99.10) from two sources. The parameters of the pterin are of the same order as those of the most effective substrates of dihydropteridine reductase. The Km values of the pterin are one order of magnitude smaller than those of the pyrimidinone, although the kcat. values are of the same order.  相似文献   

3.
The oxidation of 4,5-diaminopyrimidin-6(1H)-one, 5,6,7,8-tetrahydropteridin-4(3H)-one, its 6-methyl and cis-6,7-dimethyl derivatives, and 6-methyl- and cis-6-7-dimethyl-5,6,7,8-tetrahydropterins, by horseradish peroxidase/H2O2 is enzymic and follows Michaelis-Menten kinetics, and its Km and kcat. values were determined. This oxidation of 5,6,7,8-tetrahydropterins produces quinonoid dihydropterins of established structure, and they are known to be specific substrates for dihydropteridine reductase. By analogy the peroxidase/H2O2 oxidation of the 5,6,7,8-tetrahydropteridin-4(3H)-ones should produce similar quinonoid dihydro species. The quinonoid species derived from 5,6,7,8-tetrahydropteridin-4(3H)-one and its 6-methyl and cis-6,7-dimethyl derivatives are shown to be viable substrates for human brain dihydropteridine reductase, and apparent Km and Vmax. values are reported.  相似文献   

4.
Potassium tetrachloroplatinate (K2PtCl4) inactivates dihydropteridine reductase from human brain in a time-dependent and irreversible manner. The inactivation has been followed by measuring enzyme activity and fluorescence changes. The enzyme is completely protected from inactivation by NADH, the pterin cofactor [quinonoid 6-methyl-7,8-dihydro(6H)pterin] and dithiothreitol. Evidence is presented that K2PtCl4 reacts at the active site and that (a) thiol group(s) is involved in, or is masked by, this reaction. K2PtCl4 is a stronger inhibitor of human brain dihydropteridine reductase that cis- and trans-diaminodichloroplatinum, cis-dichloro[ethylenediamine]platinum and K4Fe(CN)6, whereas H2PtCl6 is considerably weaker and (Ph3P)3RhCl is inactive.  相似文献   

5.
Kinetic constants for the interaction of NADH and NADPH with native rat dihydropteridine reductase (DHPR) and an Escherichia coli expressed mutant (D-37-I) have been determined. Comparison of kcat and Km values measured employing quinonoid 6,7-dimethyldihydropteridine (q-PtH2) as substrate indicate that the native enzyme has a considerable preference for NADH with an optimum kcat/Km of 12 microM-1 s-1 compared with a figure of 0.25 microM-1 s-1 for NADPH. Although the mutant enzyme still displays an apparent preference for NADH (kcat/Km = 1.2 microM-1 s-1) compared with NADPH (kcat/Km = 0.6 microM-1 s-1), kinetic analysis indicates that NADH and NADPH have comparable stickiness in the D-37-I mutant. The dihydropteridine site is less affected, since the Km for q-PtH2 and K(is) for aminopterin are unchanged and the 14-26-fold synergy seen for aminopterin binding to E.NAD(P)H versus free E is decreased by less than 2-fold in the D-37-I mutant. No significant changes in log kcat and log kcat/Km versus pH profiles for NADH and NADPH were seen for the D-37-I mutant enzyme. However, the mutant enzyme is less stable to proteolytic degradation, to elevated temperature, and to increasing concentrations of urea and salt than the wild type. NADPH provides maximal protection against inactivation in all cases for both the native and D-37-I mutant enzymes. Examination of the rat DHPR sequence shows a typical dinucleotide binding fold with Asp-37 located precisely in the position predicted for the acidic residue that participates in hydrogen bond formation with the 2'-hydroxyl moiety of all known NAD-dependent dehydrogenases. This assignment is consistent with x-ray crystallographic results that localize the aspartate 37 carboxyl within ideal hydrogen bonding distance of the 2'- and 3'-hydroxyl moieties of adenosine ribose in the binary E.NADH complex.  相似文献   

6.
Kinetic studies of tetrameric recombinant human tyrosine hydroxylase isoform 1 (hTH1) have revealed properties so far not reported for this enzyme. Firstly, with the natural cofactor (6R)-Lerythro-5,6,7, 8-tetrahydrobiopterin (H4biopterin) a time-dependent change (burst) in enzyme activity was observed, with a half-time of about 20 s for the kinetic transient. Secondly, nonhyperbolic saturation behaviour was found for H4biopterin with a pronounced negative cooperativity (0.39 < h < 0.58; [S]0.5 = 24 +/- 4 microM). On phosphorylation of Ser40 by protein kinase A, the affinity for H4biopterin increased ([S]0.5 = 11 +/- 2 microM) and the negative cooperativity was amplified (h = 0.27 +/- 0.03). The dimeric C-terminal deletion mutant (Delta473-528) of hTH1 also showed negative cooperativity of H4biopterin binding (h = 0.4). Cooperativity was not observed with the cofactor analogues 6-methyl-5,6,7,8-tetrahydropterin (h = 0.9 +/- 0.1; Km = 62.7 +/- 5.7 microM) and 3-methyl-5,6,7, 8-tetrahydropterin (H43-methyl-pterin)(h = 1.0 +/- 0.1; Km = 687 +/- 50 microM). In the presence of 1 mM H43-methyl-pterin, used as a competitive cofactor analogue to BH4, hyperbolic saturation curves were also found for H4biopterin (h = 1.0), thus confirming the genuine nature of the kinetic negative cooperativity. This cooperativity was confirmed by real-time biospecific interaction analysis by surface plasmon resonance detection. The equilibrium binding of H4biopterin to the immobilized iron-free apoenzyme results in a saturable positive resonance unit (DeltaRU) response with negative cooperativity (h = 0.52-0.56). Infrared spectroscopic studies revealed a reduced thermal stability both of the apo-and the holo-hTH1 on binding of H4biopterin and Lerythro-dihydrobiopterin (H2biopterin). Moreover, the ligand-bound forms of the enzyme also showed a decreased resistance to limited tryptic proteolysis. These findings indicate that the binding of H4biopterin at the active site induces a destabilizing conformational change in the enzyme which could be related to the observed negative cooperativity. Thus, our studies provide new insight into the regulation of TH by the concentration of H4biopterin which may have significant implications for the physiological regulation of catecholamine biosynthesis in neuroendocrine cells.  相似文献   

7.
Inactivation of tyrosine hydroxylase by reduced pterins   总被引:1,自引:0,他引:1  
Tyrosine hydroxylase [E.C. 1.14.16.2] is inactivated by incubation with its reduced pterin cofactors L-erythro-tetrahydrobiopterin, 2-amino-4-hydroxy-6-methyl-5,6,7,8-tetrahydropterin and 2-amino-4-hydroxy-6,7-dimethyl-5,6,7,8-tetrahydropterin. Each of the two diastereoisomers of L-erythro-tetrahydrobiopterin inactivates tyrosine hydroxylase but the natural (6R) form is much more potent than the unnatural (6S) form at equimolar concentrations. The pterin analog 6-methyl-5-deazatetrahydropterin, which has no cofactor activity, also inactivates the enzyme whereas the oxidized pterins 7,8 dihydrobiopterin and biopterin do not. The inactivation process is both temperature and time dependent and results in a reduction of the Vmax for both tetrahydrobiopterin and tyrosine. Neither tyrosine nor oxygen inactivates tyrosine hydroxylase.  相似文献   

8.
The oxidation of tetrahydropterin with ferri-cytochrome c was studied using a tetrahydropterin-generating system composed of dihydropteridine reductase [EC 1.6.99.7] and NADH. Under aerobic conditions, 1.5 to 1.8 mol of cytochrome c was reduced per mol of NADH, whereas 2 mol of cytochrome c was reduced under anaerobic conditions. When superoxide dismutase [EC 1.15.1.1] was added to the system under aerobic conditions, only 1 mol of cytochrome c was reduced per mol of NADH, while the pterin oxidation was scarcely affected. Based on these results, we propose that the oxidation of tetrahydropterin to quinonoid dihydropterin proceeds via two steps: tetrahydropterin is first oxidized by ferri-cytochrome c to give a pterin intermediate, which has lost one electron, then in turn this reduces O2 to form O2-.  相似文献   

9.
D C Pike  M T Hora  S W Bailey  J E Ayling 《Biochemistry》1986,25(17):4762-4771
Homologues of 6-methyl-7,8-dihydropterin (6-Me-7,8-PH2) and 6-methyl-5,6,7,8-tetrahydropterin (6-Me-PH4), expanded in the pyrazine ring, were synthesized to determine the effect of increased strain on the chemical and enzymatic properties of the pyrimidodiazepine series. 2-Amino-4-keto-6-methyl-7,8-dihydro-3H,9H-pyrimido[4,5-b] [1,4]diazepine (6-Me-7,8-PDH2) was found to be more unstable in neutral solution than 6-Me-7,8-PH2. Its decomposition appears to proceed by hydrolytic ring opening of the 5,6-imine bond, followed by autooxidation. 6-Me-7,8-PDH2 can be reduced, either chemically or by dihydrofolate reductase (Km = 0.16 mM), to the 5,6,7,8-tetrahydro form (6-Me-PDH4). This can be oxidized with halogen to quinoid dihydropyrimidodiazepine (quinoid 6-Me-PDH2), which is a substrate for dihydropteridine reductase (Km = 33 microM). Whereas quinoid 6-methyldihydropterin was found to tautomerize to 6-Me-7,8-PH2 in 95% yield in 0.1 M tris(hydroxymethyl)aminomethane hydrochloride (Tris-HCl), pH 7.4, quinoid 6-Me-PDH2 gives only 53% 6-Me-7,8-PDH2, the remainder decomposing via an initial opening of the diazepine ring. Additional evidence for the extra strain in the pyrimidodiazepine system is the cyclization of quinoid 6-N-(2'-aminopropyl)divicine to quinoid 6-Me-PH2 in 57% yield in 0.1 M Tris-HCl, pH 7.4. By comparison, no quinoid 6-Me-PDH2 is formed from the homologue quinoid 6-N-(3'-aminobutyl)divicine. A small (2%) yield of 6-Me-PDH4 is found if the unstable C4a-carbinolamine intermediate is trapped by enzymatic dehydration and reduction. Although phenylalanine hydroxylase utilizes 6-Me-PDH4 (Km = 0.15 mM), the maximum velocity of tyrosine production is 20 times slower than that with 6-Me-PH4, indicating that a ring opening reaction is not a rate-limiting step in the hydroxylase pathway. Further, the maximum velocities of 2,5,6-triamino-4(3H)-pyrimidinone, 2,6-diamino-5-(methylamino)-4(3H)-pyrimidinone, and 2,6-diamino-5-(benzylamino)-4(3H)-pyrimidinone span a 35-fold range. These cofactors would theoretically form the same oxide of quinoid divicine if oxygen activation involves a carbonyl oxide intermediate. Thus, the limiting step is also not transfer of oxygen from this hypothetical intermediate to the phenylalanine substrate.  相似文献   

10.
Samples of quinonoid-l -erythrodihydrobiopterin (q-BH2) and quinonoid-6-methyl-dihydro-pterin (q-6-MPH2) were prepared by oxidation of l -erythro-5,6,7,8-tetrahydrobiopterin (BH4) and 5,6,7,8-tetrahydro-6-methylpterin (6-MPH4) and separated from D-erythro-7,8-dihydrobiopterin (7,8-BH2) and 6-methyl-7,8-dihydropterin (7,8-6-MPH2) as well as from the tetrahydropterins on phosphocellulose column by high-pressure liquid chromatography. The quinonoid dihydropterins were identified and quantitated by scan of their ultraviolet absorption and fluorescence emission spectra through their rearrangement to their 7,8-tautomer and also by gas chromatography of their rapidly synthesized trimethylsilyl derivative. Identification was also achieved by the enzymatic reduction of [3H]q-BH2to [3H]BH4 by dihydrofolate reductase (DHFR). Direct proof for the enzymatic synthesis of the q-BH2 from GTP or from 2-amino-6-(5′-triphosphoribosyl)-amino-5- or -6-formamido-6-hydroxypyrimi-dine (FPyd-P3) was obtained by isolation of the compound which was identical in all respects to the q-BH2 obtained by chemical synthesis from BH4. The reduction of enzymatically synthesized q-BH2 by dihydropteridine reductase (DHPR) to BH4 was not inhibited by methotrexate (MTX). When the enzymatically synthesized q-BH2 was converted to 7,8-BH2, it was reduced only by DHFR. This reduction, however, was inhibited by MTX. On the biosynthetic pathway from GTP to dihydrobiopterin, the enzyme responsible for the appearance of the quinonoid structure is the d -erythro-dihydroneopterin triphosphate synthetase, the product of which (quinonoid d -erythro-dihydroneopterin triphosphate) is converted to quinonoid dihydrobiopterin by l -erythro-dihydrobiopterin synthetase. Experiments in vivo established that DHFR does not participate in the reduction of dihydrobiopterin to tetra-hydrobiopterin when the former is synthesized from GTP de novo. MTX at 5 × 10?6M exerted no inhibition on the reduction of the biosynthetic dihydrobiopterin to tetrahydrobiopterin in vivo, yet completely inhibited the reduction of intraventricularly injected tritiated dihydrofolate ([3H]FH2) to tritiated tetrahydrofolate ([3H]FH4).  相似文献   

11.
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   

12.
The relationship between the NADH:lipoamide reductase and NADH:quinone reductase reactions of pig heart lipoamide dehydrogenase (EC 1.6.4.3) was investigated. At pH 7.0 the catalytic constant of the quinone reductase reaction (kcat.) is 70 s-1 and the rate constant of the active-centre reduction by NADH (kcat./Km) is 9.2 x 10(5) M-1.s-1. These constants are almost an order lower than those for the lipoamide reductase reaction. The maximal quinone reductase activity is observed at pH 6.0-5.5. The use of [4(S)-2H]NADH as substrate decreases kcat./Km for the lipoamide reductase reaction and both kcat. and kcat./Km for the quinone reductase reaction. The kcat./Km values for quinones in this case are decreased 1.85-3.0-fold. NAD+ is a more effective inhibitor in the quinone reductase reaction than in the lipoamide reductase reaction. The pattern of inhibition reflects the shift of the reaction equilibrium. Various forms of the four-electron-reduced enzyme are believed to reduce quinones. Simple and 'hybrid ping-pong' mechanisms of this reaction are discussed. The logarithms of kcat./Km for quinones are hyperbolically dependent on their single-electron reduction potentials (E1(7]. A three-step mechanism for a mixed one-electron and two-electron reduction of quinones by lipoamide dehydrogenase is proposed.  相似文献   

13.
The catalytic properties of a new type of dihydropteridine reductase, NADPH-specific dihydropteridine reductase [EC 1.6.99.10], from bovine liver, were studied and compared with those of the previously characterized enzyme, NADH-specific dihydropteridine reductase [EC 1.6.99.7]. With quinonoid-dihydro-6-methylpterin, approximate Km values of NADPH-specific dihydropteridine reductase for NADPH and NADH were estimated to be 1.4 micron and 2,900 microns, respectively. The Vmax values were 1.34 mumol/min/mg with NADPH and 1.02 mumol/min/mg with NADPH. With NADPH, the Km values of the enzyme for the quinonoid-dihydro forms of 6-methylpterin and biopterin were 1.4 micron and 6.8 microns, respectively. The enzyme was inhibited by its reaction product, NADP+, in a competitive manner, and the inhibition constant was determined to be 3.2 microns. The enzyme was severely inhibited by L-thyroxine and by 2,6-dichlorophenolindophenol.  相似文献   

14.
Physarum polycephalum is one of few non-animal organisms capable of synthesizing tetrahydrobiopterin from GTP. Here we demonstrate developmentally regulated expression of quinoid dihydropteridine reductase (EC 1.6.99.7), an enzyme required for recycling 6,7-[8H]-dihydrobiopterin. Physarum also expresses phenylalanine-4-hydroxylase activity, an enzyme that depends on dihydropteridine reductase. The 24.4 kDa Physarum dihydropteridine reductase shares 43% amino acid identity with the human protein. A number of residues important for function of the mammalian enzyme are also conserved in the Physarum sequence. In comparison to sheep liver dihydropteridine reductase, purified recombinant Physarum dihydropteridine reductase prefers pterin substrates with a 6-(1', 2'-dihydroxypropyl) group. Our results demonstrate that Physarum synthesizes, utilizes and metabolizes tetrahydrobiopterin in a way hitherto thought to be restricted to the animal kingdom.  相似文献   

15.
The structure of dyspropterin, a new name given to an intermediate which is formed from dihydroneopterin triphosphate in the biosynthetic pathway of tetrahydrobiopterin, has been studied. Sepiapterin reductase (EC 1.1.1.153) was found to reduce dyspropterin to tetrahydrobiopterin in the presence of NADPH. Several lines of evidence showing the formation of tetrahydrobiopterin have been presented. Stoichiometric analysis revealed that there is a 1:2 relationship between the production of biopterin and the oxidation of NADPH during the reductase-catalyzed reduction of dyspropterin. The tetrahydrobiopterin production from dyspropterin was enhanced by dihydropteridine reductase (EC 1.6.99.7). Dyspropterin could also serve as a cofactor in phenylalanine hydroxylase (EC 1.14.16.1) system. These results are consistent with the view that dyspropterin is 6-(1,2-dioxopropyl)-5,6,7,8-tetrahydropterin. Based on our findings, the biosynthetic pathway of tetrahydrobiopterin from dihydroneopterin triphosphate has been discussed.  相似文献   

16.
A recently described new form of hyperphenylalaninemia is characterized by the excretion of 7-substituted isomers of biopterin and neopterin and 7-oxo-biopterin in the urine of patients. It has been shown that the 7-substituted isomers of biopterin and neopterin derive from L-tetrahydrobiopterin and D-tetrahydroneopterin and are formed during hydroxylation of phenylalanine to tyrosine with rat liver dehydratase-free phenylalanine hydroxylase. We have now obtained identical results using human phenylalanine hydroxylase. The identity of the pterin formed in vitro and derived from L-tetrahydrobiopterin as 7-(1',2'-dihydroxypropyl)pterin was proven by gas-chromatography mass spectrometry. Tetrahydroneopterin and 6-hydroxymethyltetrahydropterin also are converted to their corresponding 7-substituted isomers and serve as cofactors in the phenylalanine hydroxylase reaction. Dihydroneopterin is converted by dihydrofolate reductase to the tetrahydro form which is biologically active as a cofactor for the aromatic amino acid monooxygenases. The 6-substituted pterin to 7-substituted pterin conversion occurs in the absence of pterin-4a-carbinolamine dehydratase and is shown to be a nonenzymatic process. 7-Tetrahydrobiopterin is both a substrate (cofactor) and a competitive inhibitor with 6-tetrahydrobiopterin (Ki approximately 8 microM) in the phenylalanine hydroxylase reaction. For the first time, the formation of 7-substituted pterins from their 6-substituted isomers has been demonstrated with tyrosine hydroxylase, another important mammalian enzyme which functions in the hydroxylation of phenylalanine and tyrosine.  相似文献   

17.
本文报导了从猪肝中提取二氢蝶啶还原酶[Ecl.6.99.7]的方法,提取百分率达30%左右。以DMPH_4为底物,分别以NADH和NADPH为辅酶,测定了该酶的动力学,发现它对NADH具有一定的特异性[Km(NADH)Vmax(NADPH)]。不同的金属离子对该酶活性影响的程度有很大的差异。  相似文献   

18.
Tetrahydrobiopterin ((6R)-5,6,7,8-tetrahydro-L-biopterin (H4biopterin)) is an essential cofactor of nitric-oxide synthases (NOSs), but its role in enzyme function is not known. Binding of the pterin affects the electronic structure of the prosthetic heme group in the oxygenase domain and results in a pronounced stabilization of the active homodimeric structure of the protein. However, these allosteric effects are also produced by the potent pterin antagonist of NOS, 4-amino-H4biopterin, suggesting that the natural cofactor has an additional, as yet unknown catalytic function. Here we show that the 5-methyl analog of H4biopterin, which does not react with O2, is a functionally active pterin cofactor of neuronal NOS. Activation of the H4biopterin-free enzyme occurred in a biphasic manner with half-maximally effective concentrations of approximately 0.2 microM and 10 mM 5-methyl-H4biopterin. Thus, the affinity of the 5-methyl compound was 3 orders of magnitude lower than that of the natural cofactor, allowing the direct demonstration of the functional anticooperativity of the two pterin binding sites of dimeric NOS. In contrast to H4biopterin, which inactivates nitric oxide (NO) through nonenzymatic superoxide formation, up to 1 mM of the 5-methyl derivative did not consume O2 and had no effect on NO steady-state concentrations measured electrochemically with a Clark-type NO electrode. Therefore, reconstitution with 5-methyl-H4biopterin allowed, for the first time, the detection of enzymatic NO formation in the absence of superoxide or NO scavengers. These results unequivocally identify free NO as a NOS product and indicate that reductive O2 activation by the pterin cofactor is not essential to NO biosynthesis.  相似文献   

19.
Human Hageman factor, a plasma proteinase zymogen, was activated in vitro under a near physiological condition (pH 7.8, ionic strength I = 0.14, 37 degrees C) by Pseudomonas aeruginosa elastase, which is a zinc-dependent tissue destructive neutral proteinase. This activation was completely inhibited by a specific inhibitor of the elastase, HONHCOCH(CH2C6H5)CO-Ala-Gly-NH2, at a concentration as low as 10 microM. In this activation Hagemen factor was cleaved, in a limited fashion, liberating two fragments with apparent molecular masses of 40 and 30 kDa, respectively. The appearance of the latter seemed to correspond chronologically to the generation of activated Hageman factor. Kinetic parameters of the enzymatic activation were kcat = 5.8 x 10(-3) s-1, Km = 4.3 x 10(-7) M and kcat/Km = 1.4 x 10(4) M-1 x s-1. This Km value is close to the plasma concentration of Hageman factor. Another zinc-dependent proteinase, P. aeruginosa alkaline proteinase, showed a negligible Hageman factor activation. In the presence of a negatively charged soluble substance, dextran sulfate (0.3-3 micrograms/ml), the activation rate by the elastase increased several fold, with the kinetic parameters of kcat = 13.9 x 10(-3) s-1, Km = 1.6 x 10(-7) M and kcat/Km = 8.5 x 10(4) M-1 x s-1. These results suggested a participation of the Hageman factor-dependent system in the inflammatory response to pseudomonal infections, due to the initiation of the system by the bacterial elastase.  相似文献   

20.
9 partially purified enzyme (Enzyme A) from Drosophila melanogaster Aatalyzes the conversion of 7,8- dihydroneopterin triphosphate to a compound that, from its ultraviolet absorption spectrum and other characteristics, appears to be 6- pyruvoyl -tetrahydropterin. This product can be converted to 6-lactoyl-tetrahydropterin in the presence of another partially purified enzyme (Enzyme B) and NADPH, and to 5,6,7,8-tetrahydrobiopterin in the presence of a third enzyme preparation (biopterin synthase) and NADPH. The enzymically-produced 6-lactoyl-tetrahydropterin, when exposed to air, is oxidized nonenzymically to sepiapterin (6-lactoyl-7,8- dihydropterin ). The results indicate that although 6-lactoyl-tetrahydropterin can be converted enzymically to tetrahydrobiopterin, neither it nor sepiapterin is an obligate intermediate in the conversion of 7,8- dihydroneopterin triphosphate to tetrahydrobiopterin.  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