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.
Biosynthesis of methanopterin   总被引:6,自引:0,他引:6  
R H White 《Biochemistry》1990,29(22):5397-5404
The biosynthetic pathway for the generation of the methylated pterin in methanopterins was determined for the methanogenic bacteria Methanococcus volta and Methanobacterium formicicum. Extracts of M. volta were found to readily cleave L-7,8-dihydroneopterin to 7,8-dihydro-6-(hydroxymethyl)pterin, which was confirmed to be a precursor of the pterin portion of the methanopterin. [methylene-2H]-6-(Hydroxymethyl)pterin was incorporated into methanopterin by growing cells of M. volta to an extent of 30%. Both the C-11 and C-12 methyl groups of methanopterin originate from [methyl-2H3]methionine, as confirmed by the incorporation of two C2H3 groups into 6-ethyl-7-methylpterin, a pterin-containing fragment derived from methanopterin. Cells grown in the presence of [methylene-2H]-6-(hydroxymethyl)pterin, [ethyl-2H4]-6-[1 (RS)-hydroxyethyl]pterin, [methyl-2H3]-6- (hydroxymethyl)-7-methylpterin, [ethyl-2H4, methyl-2H3]-6-[1 (RS)-hydroxyethyl]-7-methylpterin, and [1-ethyl-3H]-6-[1 (RS)-hydroxyethyl]-7-methylpterin showed that only the non-7-methylated pterins were incorporated into methanopterin. Cells extracts of M. formicicum readily condensed synthetic [methylene-3H]-7,8-H2-6-(hydroxymethyl)pterin-PP with methaniline to generate demethylated methanopterin, which is then methylated to methanopterin by the cell extract in the presence of S-adenosylmethionine. These observations indicate that the pterin portion of methanopterin is biosynthetically derived from 7,8-H2-6-(hydroxymethyl)pterin, which is coupled to methaniline by a pathway analogous to the biosynthesis of folic acid. This pathway for the biosynthesis of methanopterin represents the first example of the modification of the specificity of a coenzyme through a methylation reaction.  相似文献   

2.
Biosynthesis of the 7-methylated pterin of methanopterin.   总被引:4,自引:3,他引:1       下载免费PDF全文
The incorporation of [15N]glycine and [U-methyl-2H]methionine into methanopterin by growing cells of a methanogenic bacterium was measured to establish the biosynthetic route of the methylated pterin in the structure. The tetrahydromethanopterin produced by the cells was oxidatively cleaved to produce 7-methylpterin, and the amount of label incorporated into this pterin was measured by gas chromatography-mass spectrometry of the ditrimethylsilyl derivative of this compound. Approximately 27% of the 7-methylpterin and the guanine present in the cell was derived from the fed [15N]glycine. [U-methyl-2H]methionine was incorporated with the initial retention of all three deuteriums. These results are consistent with the biosynthesis of the pterin of methanopterin originating from GTP and its 7-methyl group arising from the methyl group of methionine.  相似文献   

3.
Cofactor extracts of Methanogenium tationis were screened for the presence of pterin-derivatives. Methanopterin, sarcinapterin and 7-methylpterin were absent, while 2-amino-4-hydroxy-pteridine and another blue fluorescent compound with a pterin spectrum were detected. The latter pterin was purified by ion exchange and reversed-phase column chromatography. The structure of this compound was elucidated by combining spectrophotometry, amino acid analysis and 1H-NMR spectroscopy. The pterin, which we named tatiopterin, was identified as an aspartyl derivative of sarcinapterin with a 7-proton instead of a 7-methyl group in the pterin moiety. The IUPAC name is: N-[-1'-(2'-amino-4'-hydroxy-7'-proton-6'-pteridinyl)ethyl]-4- [2',3',4',5'-tetrahydroxypent-1'-yl(5'----1')O-alpha- ribofuranosyl-5'-phosphoric acid]aniline, in which the phosphate group is esterified with alpha-hydroxyglutarylglutamylaspartic acid.  相似文献   

4.
Analyses were made of the structures and levels of folates and modified folates present in extremely thermophilic bacteria. These procedures involved the chemical analysis of products resulting from the oxidative cleavage of the 6-substituted, folatelike tetrahydropterins present in the cells. Air-oxidized cell extracts of extreme thermophiles from two members of the archaebacterial order Thermococcales, Thermococcus celer and Pyrococcus furiosus, contained only 7-methylpterin, indicating that these cells contain a modified folate with a methylated pterin. Cell extracts also contained 6-acetyl-7-methyl-7,8-dihydropterin, another product derived from the oxidative cleavage of a dimethylated folate, demonstrating that both the C-7 and C-9 carbons of the pterin were methylated. Extracts, however, contained neither p-aminobenzoylpolyglutamates nor methaniline, the oxidative cleavage products of folates and methanopterin, respectively, indicating that they contain a previously undescribed C1 carrier(s). On the basis of the level of the 7-methylpterin isolated, the levels of modified folate were 2 to 10 times higher than those typically found in mesophilic bacteria and 10 to 100 times less than the level of methanopterin found in the methanogenic bacteria. Oxidized cell extracts of Sulfolobus spp. of the archaebacterial order Sulfolobales contained only pterin, and, like members of the order Thermococcales, they contained neither-p-aminobenzoylpolyglutamates nor methaniline. Oxidized cell extracts of the extreme thermophiles Pyrobaculum sp. strain H10 and Pyrodictium occultum, from the archaebacterial orders Thermoproteales and Pyrodictiales, respectively, and Thermotoga maritima from the eubacterial order Thermotogales, contained pterin and p-aminobenzoylpolyglutamates, indicating that these cells contained unmodified folates. The levels of p-aminobenzoylpolyglutamates in these archaebacterial cell extracts indicate that the folates were present in the cells at levels 4 to 10 times higher than generally found in those mesophilic eubacteria which do not folates in energy metabolism. The levels and chain lengths of the of p-aminobenzoylpolyglutamates present in Thermotoga maritima were typical of those found in mesophilic eubacteria.  相似文献   

5.
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   

6.
The partial characterization of the modified folate present in Sulfolobus solfataricus has been carried out. Separation of ethanol-water extracts of these cells on a DEAE-Sephadex column led to the isolation of a small amount of intact oxidized cofactor, which, when subjected to reductive cleavage with Zn-HCl, produced 6-methylpterin. This indicated that the modified folate in these cells contained a nonmethylated pterin linked, via a methylene group at the C-6 position of the pterin, to an arylamine, as is found in folate. Oxidative cleavage of intact reduced cofactor produced pterin and a single arylamine. The azo dye derivative of this arylamine was prepared and purified by chromatography on a Bio-Gel P-6 column. The resulting purified compound was shown to be readily hydrolyzed in dilute acid to the azo dye derivative of 5-(p-aminophenyl)-1,2,3,4-tetrahydroxypantane, which was, in turn, readily cleaved to 5-(p-aminophenyl)-1,2,3,4- tetrahydroxypentane by Zn-HCl reduction. The stereochemistry of the resulting 5-(p-aminophenyl)-1,2,3,4-tetrahydroxypentane was shown to be ribo, the same as that of the 5-(p-aminophenyl)-1,2,3,4- tetrahydroxypentane moiety found in methanopterin. The complete arylamine side chain of the modified folate thus contains 5-(p-aminophenyl)-1,2,3,4-tetrahydroxypentane attached, via an acid-labile bond, to a currently unidentified substituent. The modified folate present in S. solfataricus thus contains structural features common to both folates and methanopterin.  相似文献   

7.
Identification of a pterin derivative in Escherichia coli DNA photolyase   总被引:2,自引:0,他引:2  
B Y Wang  S P Jordan  M S Jorns 《Biochemistry》1988,27(12):4222-4226
DNA photolyase from Escherichia coli contains reduced flavin adenine dinucleotide plus a second chromophore, partially characterized in previous studies. Both chromophores function as sensitizers in catalysis. The second chromophore has been identified as a 6-substituted pterin derivative. The compound is oxidized with permanganate to yield 6-carboxypterin or reduced with sodium cyanoborohydride to yield a 5,6,7,8-tetrahydropterin derivative. The second chromophore exhibits spectral properties (lambda max = 360, 255 nm, pH 2) similar to that observed for 7,8-dihydropterin cations. The compound does not exhibit a spectrally detectable pKa around 4 but is converted to a dication (lambda max = 346, 255 nm) in strong acid (pKa approximately 1). Similar ionization behavior is observed with 7,8-dihydropterin derivatives that are alkylated at N(5). The instability of the second chromophore in weakly alkaline solution is due to a fully reversible conversion to a labile bleached form. As compared with other pterin derivatives, the hydrolytic instability is unusual but is very similar to that observed for 5,6-dialkyl-7,8-dihydropterinium salts. It is proposed that the second chromophore is a 7,8-dihydropterin with substituents at positions 5 and 6. The discovery that a pterin derivative functions as a photosensitizer in DNA repair is apparently the first example of a photobiological function for pterins.  相似文献   

8.
Abstract Tyrosine hydroxylase (TH, EC 1.14.16.2) from beef brain striata was purified 23-fold from an extract of an acetone powder. If this enzyme preparation is treated with a cyclic AMP-dependent protein phosphorylation system, there is a change in the pH dependence of the enzyme activity. The pH optimum at saturating tetrahydrobiopterin (BH4) concentration is shifted from below pH 6 to about pH 6.7. At pH 7, activation is expressed mainly as an increase in Vmax, whereas at pH 6, activation is expressed mainly as a decrease in Km for the pterin cofactor. Further, even with the control enzyme the Km for pterin cofactor declines precipitously as the pH is increased from 6 toward neutrality. Similar data were obtained with G-25 Sephadex-treated rat striatal TH. Experiments in which rat striatal synaptosomes were used demonstrated that the in situ activation of TH by phosphorylating conditions is expressed primarily as an increase in the maximum rate of dopamine synthesis. These results indicate that changes in TH activity caused by cyclic AMP-dependent protein phosphorylation will depend to a large extent on the pH of the TH environment.  相似文献   

9.
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   

10.
Tyrosine 3-monooxygenase activity of the crude extract from rat striatum had a sharp pH optimum at pH 5.4 and showed almost no activity at or above pH 7. When the crude extract was partially purified by pH precipitation and chromatography on DEAE-cellulose, the enzyme showed a high activity in the pH range of 5.8 to 7.4. Incubation of the partially purified enzyme with catecholamines such as dopamine, norepinephrine, and epinephrine resulted in a remarkable decrease in the enzyme activity, as assayed at a neutral pH. This suppression of the enzyme activity by catecholamines differed from the well-known feedback inhibition which is competitive with respect to the pterin cofactor; the former occurred at a very much lower concentration of catecholamines even in the presence of a near-saturating concentration of a pterin cofactor, and the former was a time-dependent reaction. The enzyme, the activity of which had been suppressed by the incubation with dopamine, was remarkably activated by the incubation with the catalytic subunit of cyclic AMP-dependent protein kinase in the presence of an ATP-generating system. These results suggest that the activity of tyrosine 3-monooxygenase may be suppressed by its end products in a normal state and it may be stimulated by cyclic AMP-dependent protein kinase as occasion demands.  相似文献   

11.
Queuine, a modified form of 7-deazaguanine present in certain transfer RNAs, is shown to occur in Drosophila melanogaster adults in a free form and its concentration varies as a function of age, nutrition and genotype. In several, but not all mutant strains, the concentrations of queuine and the Q(+) (queuine-containing) form of tRNATyr are correlated. The bioassay employs L-M cells which respond to the presence of queuine by an increase in their Q(+)tRNAAsp that is accompanied by a decrease in the Q(-)tRNAAsp isoacceptors. The increase in Q(+)tRNATyr in Drosophila that occurs on a yeast diet is accompanied by an increase in queuine. Similarly the increase of Q(+)tRNAs with age also is accompanied by an increase in free queuine. In two mutants, brown and sepia, these correlations were either diminished or failed to occur. Indeed, the extract of both mutants inhibited the response of the L-M cells to authentic queuine. When the pteridines that occur at abnormally high levels in sepia were used at 1 x 10(-6)M, the inhibition of the L-M cell assay occurred in the order biopterin greater than pterin greater than sepiapterin. These pteridines were also inhibitory for the purified guanine:tRNA transglycosylase from rabbit but the relative effectiveness then was pterin greater than biopterin greater than sepiapterin. Pterin was competitive with guanine in the enzyme reaction with Ki = 0.9 x 10(-7)M. Also when an extract of sepia was chromatographed on Sephadex G-50, the pteridine-containing fractions only were inhibitory toward the L-M cell assay or the enzyme assay. These results indicate that free queuine occurs in Drosophila but also that certain pteridines may interfere with the incorporation of queuine into RNA.  相似文献   

12.
Inducible nitric-oxide synthase (NOS) was expressed and purified in the absence of 6(R)-tetrahydro-l-biopterin (H(4)B). Pterin-free NOS exhibits a Soret band (416-420 nm) characteristic of predominantly low spin heme and does not catalyze the formation of nitric oxide (. NO) (Rusche, K. M., Spiering, M. M., and Marletta, M. A. (1998) Biochemistry 37, 15503-15512). Reconstitution of pterin-free NOS with H(4)B was monitored by a shift in the Soret band to 396-400 nm, the recovery of.NO-forming activity, and the measurement of H(4)B bound to the enzyme. As assessed by these properties, H(4)B binding was not rapid and required the presence of a reduced thiol. Spectral changes and recovery of activity were incomplete in the absence of reduced thiol. Full reconstitution of holoenzyme activity and stoichiometric H(4)B binding was achieved in the presence of 5 mm glutathione (GSH). Preincubation with GSH before the addition of H(4)B decreased, whereas lower concentrations of GSH extended, the time required for reconstitution. Six protected cysteine residues in pterin-free NOS were identified by labeling of NOS with cysteine-directed reagents before and after reduction with GSH. Heme and metal content of pterin-free and H(4)B-reconstituted NOS were also measured and were found to be independent of H(4)B content. Additionally, pterin-free NOS was reconstituted with 6-methylpterin analogs, including redox-stable deazapterins. Reconstitution with the redox-stable pterin analogs was neither time- nor thiol-dependent. Apparent binding constants were determined for the 6-methyl- (50 microm) and 6-ethoxymethyl (200 microm) deazapterins. The redox-stable pterin analogs appear to bind to NOS in a different manner than H(4)B.  相似文献   

13.
J M Hevel  M A Marletta 《Biochemistry》1992,31(31):7160-7165
Nitric oxide synthase (NOS) (EC 1.14.23) catalyzes the oxidation of L-arginine to citrulline and nitric oxide. The complex reaction carried out by NOS, which involves NADPH, O2, and enzyme-bound FAD, FMN, and tetrahydrobiopterin (BH4), has only recently begun to be elucidated. Herein we report the characterization of the pterin requirement of murine macrophage NOS. Although purified NOS activity was not dependent on BH4, activity was significantly enhanced by BH4 in a concentration-dependent fashion. NOS purified in the absence of added BH4 was found to contain substoichiometric concentrations of enzyme-bound pterin, where increased concentrations of bound pterin correlated with an increase in activity when assayed in the absence of exogenous BH4. However, NOS purified in the presence of BH4 followed by gel filtration exhibited a 1 mol of pterin:1 mol of NOS 130-kDa subunit stoichiometry and activity that was essentially independent of exogenous BH4. Experiments to probe a redox role for the pterin were carried out using pterin analogues. 6(R,S)-Methyltetrahydropterin was found to increase NOS activity in enzyme purified in the absence of BH4. However, the deaza analogue, 6(R,S)-methyl-5-deazatetrahydropterin, was not only incapable of supporting enzymatic turnover but also inhibited citrulline formation in a concentration-dependent manner. Overall, these results support a role for BH4 in the NOS reaction that involves stabilization of the enzyme and redox chemistry wherein a 1:1 stoichiometry between bound pterin and NOS subunit results in maximum activity.  相似文献   

14.
7-Methylpterin and 7-methyllumizine were isolated and identified in extracts of methanogenic bacteria which had been extracted in air with ethanol-water. Ethanol-water preparations of cells extracted under nitrogen or hydrogen were devoid of these compounds. Extracts of cells obtained in the presence of air also had an increased amount of a complex arylamine which, on acid hydrolysis, gave 1 mol each of phosphate, 5-(p-aminophenyl)-1,2,3,4-tetrahydroxypentane, and alpha-hydroxyglutaric acid. Gas chromatography-mass spectrometry was used to identify the 5-(p-aminophenyl)-1,2,3,4-tetrahydroxypentane as its tetratrimethylsilyl derivative and the alpha-hydroxyglutaric acid as the n-butyl ester derivative of its gamma-lactone. When exposed to air, extracts of cells prepared in the absence of air produced 6-acetyl-7-methylpterin and 7-methylxanthopterin in addition to 7-methylpterin and 7-methyllumizine. It is concluded that these compounds are derived from the oxidative cleavage of the tetrahydromethanopterin, which is present in these bacteria, by a series of reactions analogous to those known to occur in the oxidative cleavage of tetrahydrofolic acid.  相似文献   

15.
A major pterin was isolated by reverse-phase high-performance liquid chromatography from cellular extract of vegetative cells of Dictyostelium discoideum after perchloric deproteinization and oxidation with acidic iodine. This compound was characterized by its chromatographic behavior, its absorption and fluorescence properties, by its oxidation product with alkaline permanganate, by secondary ion mass spectrometry and by circular dichroism. The final identification was obtained by comparison with authentic materials. It is concluded that the major pterin of D. discoideum is the compound 6-(D-threo-1,2-dihydroxypropyl)-pterin. The name dictyopterin is proposed for this new natural isomer of L-biopterin.  相似文献   

16.
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   

17.
Abstract A blue fluorescent compound was extracted and purified from cells of Methanobacterium thermoautotrophicum . The compound was identified as 7-methylpterin on the basis of its (physico-) chemical properties and by comparison with 7-methylpterin prepared by organic synthesis. The compound is present in all methanogenic bacteria studied so far and it provides methanogenic bacteria the characteristic blue fluorescence observed upon fluorescence microscopy.  相似文献   

18.
Robert H. White 《Chirality》1996,8(4):332-340
The configuration at the C-9 of methanopterin (MPT) has been determined by comparing the circular dichroism (CD) spectra of MPT and its hydrolytic fragment, 1-[4-[[1-(2-amino-7-methyl-4-hydroxy-6-pteridinyl)-ethyl]amino]phenyl]-1-deoxy-D -ribitol (HP-1), with the CD spectra of a series of model compounds of known stereochemistry. These compounds included (S)-6-[1-(4-carboxymethylanilino)ethyl]pterin, (S-6(1-hydroxyethyl)-7-methylpterin, (S-6-(1-hydroxyethyl)pterin, (R)-6-(1-phenoxyethyl)pterin, D (+)-neopterin, and L -biopterin. From this comparison it was concluded that MPT has the R configuration at C-9 and is thus configurationally related to D (+)-neopterin, which has the S configuration at C-1. From previous work establishing the relative stereochemistry at C-6, C-7, and C-9 of N5-N10-methenyl-5,6,7,8-tetrahydromethanopterin (N5-N10-methenyl-H4MPT) as R, S, and R, respectively, it is clear that the remaining asymmetric carbons at C-6 and C-7 of H4MPT have the S and S configuration, respectively. Comparison of these latter two positions to the equivalent carbons in 5,6,7,8-tetrahydrofolate (H4folate) show that the steps involved in the biological reduction of MPT to H4MPT occur with the same stereochemical outcome as those involved in the biological reduction of folate to H4folate. © 1996 Wiley-Liss, Inc.  相似文献   

19.
B Wurster  F Bek    U Butz 《Journal of bacteriology》1981,148(1):183-192
Kinetic data obtained for deamination of pterin by the extracellular fraction from Dictyostelium discoideum yielded apparently linear Lineweaver-Burk plots for pterin. The Michaelis constant for pterin was 30 microM. The data for folic acid deamination yielded convex Lineweaver-Burk plots. Convex Lineweaver-Burk plots could result from the presence of two types of enzymes with different affinities. The data for folic acid deamination were analyzed mathematically for two types of enzymes. This analysis produced Michaelis constants for folic acid of 1.8 and 23 microM competition studies suggested that an enzyme with low affinity nonspecifically catalyzed the deamination of folic acid and pterin, whereas an enzyme with high affinity was a specific folic acid deaminase. A specific folic acid deaminase with high affinity appeared to be present on the surface of D. discoideum cells. The Michaelis constant for this enzyme was 2.6 microM. Cells growing in nutrient broth and cells starved in phosphate buffer released folic acid and pterin deaminases. The quantity of deaminase activities released by the cells appeared to be controlled by chemoattractants. Starving cells that were supplied with folic acid, pterin, or adenosine 3',5'-phosphate increased their extracellular folic acid and pterin deaminase activities to a larger extent than did cell suspensions to which no chemoattractants were added. Administration of folic acid or pterin to starving cells caused increases of the activity of extracellular adenosine 3',5'-phosphate phosphodiesterase and repressed increases of the activity of phosphodiesterase inhibitor.  相似文献   

20.
In the present study the optical biosensor technique, based on the surface plasmon resonance (SPR) phenomenon, was used for real-time measurements of the reversible binding of the pterin cofactor (6R)-l-erythro-5,6,7,8-tetrahydrobiopterin (BH(4)) and l-phenylalanine (l-Phe) to human phenylalanine hydroxylase (hPAH). When BH(4) (241 Da) was injected over the sensor chip with immobilized tetrameric wt-hPAH a positive DeltaRU response was observed with a square-wave type of sensorgram and a saturable response (about 25 RU/(pmol subunit/mm(2)) with a [S](0.5) value of 5.6 +/- 0.8 microM for the pterin cofactor. The rapid on-and-off rates were, however, not possible to determine. By contrast, when l-Phe (165 Da) was injected a time-dependent increase in RU (up to about 3 min) and a much higher saturable DeltaRU response (about 75 RU/(pmol subunit/mm(2)) at 2 mM l-Phe) than expected (i.e., <5 RU/(pmol subunit/mm(2))) from the low molecular mass of l-Phe were observed in the sensorgram. The half-time for the on-and-off rates were 6 +/- 2 and 9 +/- 1 s, respectively, at 2 mM l-Phe. The steady-state (apparent equilibrium) response revealed a hyperbolic concentration dependence with a [S](0.5) value of 98 +/- 7 microM. The [S](0.5) values of both pterin cofactor and l-Phe were lower than those determined by steady-state enzyme kinetic analysis. Evidence is presented that the DeltaRU response to l-Phe is accounted for by the global conformational transition which occurs in the enzyme upon l-Phe binding, i.e., by the slow reversible transition from a low activity state ("T"-state) to a high activity state ("R"-state) characteristic of this hysteretic enzyme.  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