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.
T Toraya  A Ishida 《Biochemistry》1988,27(20):7677-7681
Cleavage of the C-Co bond of sterically hindered alkylcobalamins bearing neither an adenine moiety nor functional groups, such as isobutylcobalamin, neopentylcobalamin, and cyclohexylcobalamin, was markedly accelerated by their interaction with apoprotein of diol dehydrase, although these cobalamins do not function as coenzyme. Acceleration of the conversion of alkylcobalamins to enzyme-bound hydroxocobalamin was stoichiometric and obeyed first-order reaction kinetics. These results, together with strong competitive inhibition by these alkylcobalamins with respect to adenosylcobalamin, indicate that acceleration of the C-Co bond cleavage by the apoenzyme is due to labilization of their C-Co bond by binding to the active site of the enzyme. This labilization is considered to be caused by a steric distortion of the corrin ring which is induced by specific tight interaction of the cobalamin moiety with apoprotein. The importance of such a labilizing effect for activation of the C-Co bond of adenosylcobalamin in enzymatic reactions is discussed.  相似文献   

2.
Magnusson OT  Frey PA 《Biochemistry》2002,41(5):1695-1702
3',4'-Anhydroadenosylcobalamin (anAdoCbl) is an analogue of the adenosylcobalamin (AdoCbl) coenzyme (Magnusson, O.Th., and Frey, P. A. (2000) J. Am. Chem. Soc. 122, 8807-8813). This compound supports activity for diol dehydrase at 0.02% of that observed with AdoCbl. In a side reaction, however, anAdoCbl induces suicide inactivation by an electron-transfer mechanism. Homolytic cleavage of the Co-C bond of anAdoCbl at the active site of diol dehydrase was observed by spectrophotometric detection of cob(II)alamin. Anaerobic conversion of enzyme bound cob(II)alamin to cob(III)alamin, both in the absence and presence of substrate, indicates that the coenzyme derived 5'-deoxy-3',4'-anhydroadenosine-5'-yl serves as the oxidizing agent. This hypothesis is supported by the stoichiometric formation of 3',5'-dideoxyadenosine-4',5'-ene as the nucleoside cleavage product, as determined by high-performance liquid chromatography, mass spectrometry, and nuclear magnetic resonance spectroscopy. Experiments performed in deuterium oxide show that a single solvent exchangeable proton is incorporated into the product. These data are consistent with the intermediate formation of a transient allylic anion formed after one electron transfer from cob(II)alamin to the allylic 5'-deoxy-3',4'-anhydroadenosyl radical. Selective protonation at C3' was demonstrated by spectroscopic characterization of the purified product. This study provides an example of suicide inactivation of a radical enzyme brought about by a side reaction of an analogue of the radical intermediate.  相似文献   

3.
Coenzyme B(12) dependent diol dehydratase undergoes mechanism-based inactivation by glycerol, accompanying the irreversible cleavage of the coenzyme Co-C bond. Bachovchin et al. [Biochemistry16, 1082-1092 (1977)] reported that glycerol bound in the G(S) conformation, in which the pro-S-CH(2) OH group is oriented to the hydrogen-abstracting site, primarily contributes to the inactivation reaction. To understand the mechanism of inactivation by glycerol, we analyzed the X-ray structure of diol dehydratase complexed with cyanocobalamin and glycerol. Glycerol is bound to the active site preferentially in the same conformation as that of (S)-1,2-propanediol, i.e. in the G(S) conformation, with its 3-OH group hydrogen bonded to Serα301, but not to nearby Glnα336. k(inact) of the Sα301A, Qα336A and Sα301A/Qα336A mutants with glycerol was much smaller than that of the wild-type enzyme. k(cat) /k(inact) showed that the Sα301A and Qα336A mutants are substantially more resistant to glycerol inactivation than the wild-type enzyme, suggesting that Serα301 and Glnα336 are directly or indirectly involved in the inactivation. The degree of preference for (S)-1,2-propanediol decreased on these mutations. The substrate activities towards longer chain 1,2-diols significantly increased on the Sα301A/Qα336A double mutation, probably because these amino acid substitutions yield more space for accommodating a longer alkyl group on C3 of 1,2-diols. Database Structural data are available in the Protein Data Bank under the accession number 3AUJ. Structured digital abstract ? Diol dehydrase gamma subunit, Diol dehydrase beta subunit and Diol dehydrase alpha subunit physically interact by X-ray crystallography (View interaction).  相似文献   

4.
A triplet spin system (S=1) is detected by low-temperature electron paramagnetic resonance (EPR) spectroscopy in samples of diol dehydrase and the functional adenosylcobalamin (AdoCbl) analogue 5'-deoxy-3',4'-anhydroadenosylcobalamin (anAdoCbl). Different spectra are observed in the presence and absence of the substrate (R,S)-1,2-propanediol. In both cases, the spectra include a prominent half-field transition (DeltaM(S) = 2) that is a hallmark of strongly coupled triplet spin systems. The appearance of 59Co hyperfine splitting in the EPR signals and the positions (g values) of the signals in the spectra show that half of the triplet spin is contributed by the low-spin Co2+ of cob(II)alamin. Line width effects from isotopic labeling (13C and 2H) in the 5'-deoxy-3',4'-anhydroribosyl ring demonstrate that the other half of the spin triplet is from an allylic 5'-deoxy-3',4'-anhydroadenosyl (anhydroadenosyl) radical. The zero-field splitting (ZFS) tensors describing the magnetic dipole-dipole interactions of the component spins of the triplets have rhombic symmetry because of electron spin delocalization within the organic radical component and the proximity of the radical to the low-spin Co2+. The dipole-dipole interaction was modeled as a summation of point-dipole interactions involving the spin-bearing orbitals of the anhydroadenosyl radical and cob(II)alamin. Geometries which are consistent with the ZFS tensors in the presence and absence of the substrate position the 5'-carbon of the anhydroadenosyl radical 3.5 and 4.1 A from Co2+, respectively. Homolytic cleavage of the cobalt-carbon bond of the analogue in the absence of the substrate indicates that, in diol dehydrase, binding of the coenzyme to the protein weakens the bond prior to binding of the substrate.  相似文献   

5.
Isotope effects of 2.5, 2.1, and 1.0 were measured on the conversion of [3'-3H]ADP, [3'-H]UDP, and [5-3H] UDP to the corresponding 2'-deoxynucleotides by herpes simplex virus type 1 ribonucleotide reductase. These results indicate that the reduction of either purine or pyrimidine nucleotides requires cleavage of the 3' carbon-hydrogen bond of the substrate. The substrate analogs 2'-chloro-2'-deoxyuridine 5'-diphosphate (ClUDP), 2'-deoxy-2'-fluorouridine 5'-diphosphate, and 2'-azido-2'-deoxyuridine 5'-diphosphate were time-dependent inactivators of the herpes simplex virus type 1 ribonucleotide reductase. Incubation of [3'-3H]ClUDP with the enzyme was accompanied by time-dependent release of 3H to the solvent. Reaction of [beta-32P]ClUDP with the reductase resulted in the production of inorganic pyrophosphate. These results are consistent with the enzyme-mediated cleavage of the 3' carbon-hydrogen bond of ClUDP and the subsequent conversion of the nucleotide to 2-methylene-3(2H)furanone, as previously reported with the Escherichia coli ribonucleotide reductase (Harris, G., Ator, M., and Stubbe, J. A. (1984) Biochemistry 23, 5214-5225; Ator, M., and Stubbe, J. A. (1985) Biochemistry 24, 7214-7221).  相似文献   

6.
V Pasupuleti  R Lobo  R Horton 《Steroids》1988,51(3-4):269-282
Sexual skin biopsies from 13 normal women were obtained and minces/3-h studied after adding either [3H]dihydrotestosterone (DHT) or [3H]androstanediol (3 alpha diol) to RPMI-1640 medium in a Dubnoff apparatus. Unconjugated or conjugated androgens (after hydrolysis) were purified by three chromatography steps. Formation of 3 alpha diol and 3 alpha diol glucuronide (3 alpha diolG) was linear with time. The conversion of DHT to DHT17 beta G was only 4.4 +/- 0.5%/200 mg/3 h, while conversion to 3 alpha diol was 32 +/- 1.7%. The back conversion of 3 alpha diol to DHT was 30 +/- 3% and conversion to 3 alpha diolG was 4.5 +/- 1.25%. The product of the conversion separately measured of DHT to 3 alpha diol and 3 alpha diol to 3 alpha diolG was 1.5%, which is not very different than the overall conversion rate of DHT to 3 alpha diolG of 1.4%. This study indicates that the predominant path in this tissue is DHT in equilibrium 3 alpha diol----3 alpha diolG, rather than formation of DHT17 beta G and then 3 alpha reduction to 3 alpha diolG.  相似文献   

7.
(±)-7β,8α-Dihydroxy-9β,10β-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (diol epoxide-1) and (±)-7β,8α-dihydroxy-9α,10α-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (diol epoxide-2) are highly mutagenic diol epoxide diastereomers that are formed during metabolism of the carcinogen (±)-trans-7,8-dihydroxy-7,8-dihydrobenzo[a]pyrene. Remarkable stereoselectivity has been observed on metabolism of the optically pure (+)- and (?)-enantiomers of the dihydrodiol which are obtained by separation of the diastereomeric diesters with (?)-α-methoxy-α-trifluoromethylphenylacetic acid. The high stereoselectivity in the formation of diol epoxide-1 relative to diol epoxide-2 was observed with liver microsomes from 3-methylcholanthrene-treated rats and with a purified cytochrome P-448-containing monoxygenase system where the (?)-enantiomer produced a diol epoxide-2 to diol epoxide-1 ratio of 6 : 1 and the (+)-enantiomer produced a ratio of 1 : 22. Microsomes from control and phenobarbital-treated rats were less stereospecific in the metabolism of enantiomers of BP 7,8-dihydrodiol. The ratio of diol epoxide-2 to diol epoxide-1 formed from the (?)- and (+)-enantiomers with microsomes from control rats was 2 : 1 and 1 : 6, respectively. Both enantiomers of BP 7,8-dihydrodiol were also metabolized to a phenolic derivative, tentatively identified as 6,7,8-trihydroxy-7,8-dihydrobenzo[a]pyrene, which accounted for ~30% of the total metabolites formed by microsomes from control and phenobarbital-pretreated rats whereas this metabolite represents ~5% of the total metabolites with microsomes from 3-methylcholanthrene-treated rats. With benzo[a]pyrene as substrate, liver microsomes produced the 4,5-, 7,8- and 9,10-dihydrodiol with high optical purity (>85%), and diol epoxides were also formed. Most of the optical activity in the BP 7,8-dihydrodiol was due to metabolism by the monoxygenase system rather than by epoxide hydrase, since hydration of (±)-benzo[a]pyrene 7,8-oxide by liver microsomes produced dihydrodiol which was only 8% optically pure. Thus, the stereospecificity of both the monoxygenase system and, to a lesser extent, epoxide hydrase plays important roles in the metabolic activation of benzo[a]pyrene to carcinogens and mutagens.  相似文献   

8.
X C Wang  J P Burbach 《FEBS letters》1986,197(1-2):164-168
[Cyt6]AVP-(3-9), an intermediate in the metabolism of AVP-(1-9) in vitro, was used to investigate the mechanism of formation of centrally active AVP metabolites. Exposure of [Cyt6]AVP-(3-9) to rat brain membranes resulted in formation of [Cyt6]AVP-(4-9), -(5-9), [pGlu4, Cyt6]AVP-(4-9) and AVP-(3-5), which were isolated and chemically identified. Products derived from cleavage of the C-terminus of the substrate were absent. Time-course experiments further indicated that the conversion process is predominantly mediby an aminopeptidase-like mechanism. The conversion of [Cyt6]AVP-(3-9) by membranes from hippocampus, amygdala and septum was quantitatively and qualitatively similar. The results point to a major role of aminopeptidase activity in the metabolic conversion of AVP and the formation of centrally active AVP metabolites.  相似文献   

9.
The complex trans-bis(dimethylsulfoxide)chloromethylplatinum(II) (1) is fairly soluble in water, where it undergoes multiple equilibria involving the formation of geometrically distinct [Pt(H(2)O)(DMSO)Cl(CH(3))] aqua-species. On reacting an aqueous solution of 1 with monodentate nitrogen donor ligands L, such as pyridines or amines, two well distinct patterns of behavior can be recognized: (i) a single stage fast substitution of one DMSO by the entering ligand, yielding a complex of the type trans(C,N)-[Pt(DMSO)(L)Cl(CH(3))] which contains four different groups coordinated to the metal and which undergoes a slow conversion into its cis-isomer, (ii) a double substitution affording cationic complex ions of the type cis-[Pt(L)(2)(DMSO)(CH(3))](+). When this latter reaction is carried out using sterically hindered ligands, slow rotation of the bulk ligand around the Pt[bond]N bond allows for the identification of head-to-head and head-to-tail rotamers in solution, through (1)H NMR spectrometry. The addition of chloride anion to 1 leads to the anionic species cis-[Pt(DMSO)Cl(2)(CH(3))](-), where a molecule of DMSO still remains coordinated to the metal center, despite its quite fast rate of ligand exchange (k(exch) with free DMSO=12+/-1 s(-1)). The reaction of complex 1 with bidentate ligands, such as ethylenediamine (en) or simple amino acids, leads to the cationic species [Pt(en)(DMSO)(CH(3))](+) or to the neutral [Pt(DMSO)(N[bond]O)(CH(3))], (where N[bond]-O[double bond]GlyO(-), AlaO(-)).  相似文献   

10.
Splitting of the amino proton signals of 9-ethyladenine derivatives was found in proton nuclear magnetic resonance spectra at low temperature (ca. -30 degrees C). One of the separated signals corresponds to the syn amino proton relative to the N(1) nitrogen in the adenine ring and the other to the anti one. The phenomenon is ascribable to slowing down of the hindered rotation around the N(6)-C(6) bond, which has partial double bond character. On the addition of 1-cyclohexyluracil derivatives, one of the separated signals shifts downfield. From the analysis of the concentration dependence of the signals we could estimate the population of two kinds of adenine-uracil (AU) dimers that employ the syn and anti protons, respectively. i.e. the Watson-Crick-type and the Hoogsteen-type dimers. Independent of the substitution on the uracil ring, the Hoogsteen type is predominant at 70% and the Watson-Crick type at 30% (at -56 degrees C). On the other hand, with mixtures of general kinds of 9-ethyladenine derivatives with 1-cyclohexyluracil. the substituents on the adenine ring cause the population to deviate to extreme values; i.e., either the Watson-Crick-type or the Hoogsteen-type dimer predominates. 2-Chloro-9-ethyladenine and N2-(dimethylamino)-9-ethyladenine take almost completely the Hoogsteen-type dimers, while 8-bromo-9-ethyladenine, N2-(methylamino)-9-ethyladenine, and 2-amino-9-ethylpurine predominant in the Watson-Crick-type dimers.  相似文献   

11.
Lignin peroxidase catalyses the H(2)O(2)-induced oxidation of 4-methoxybenzyltrimethylsilane by an electron transfer mechanism. The intermediate radical cation undergoes preferentially C(alpha)[bond]H deprotonation to give 4-methoxybenzaldehyde whereas C(alpha)[bond]Si bond cleavage is a minor fragmentation pathway and leads to 4-methoxybenzyl alcohol. Similar results are obtained in the oxidation catalysed by the water soluble model compound 5,10,15,20-tetra(N-methyl-4-pyridyl)porphyrinatoiron(III) pentachloride. Instead, in the oxidation promoted by the genuine one-electron transfer oxidant potassium dodecatungstocobalt(III)ate C(alpha)[bond]Si bond cleavage is the exclusive fragmentation process of the intermediate radical cation. It is suggested that in the enzymatic and biomimetic oxidations of 4-methoxybenzyltrimethylsilane the deprotonation of the intermediate radical cation is promoted by the reduced form [PorFe(IV)[double bond]O] of the active oxidant, which is an iron-oxo porphyrin radical cation.  相似文献   

12.
The pathway of LH-RH degradation by two subcellular fractions (a soluble fraction and a 25 000 X g particulate fraction) of rat hypothalamus, pituitary and cerebral cortex has been studied using high performance liquid chromatography and amino acid analysis to identify the breakdown products. The primary cleavage point in the Tyr5-Gly6 bond giving [1-5] LH-RH and [6-10] LH-RH. In the presence of dithiothreitol, cleavage of LH-RH also occurred at the Pro9-Gly10 bond giving [1-9] LH-RH. The fragment [1-5] LH-RH is further degraded sequentially from the C-terminus and [1-4] LH-RH, [1-3] LH-RH, tyrosine and tryptophan were identified. The other major fragment, [6-10] LH-RH, is rapidly broken down, the only intermediate product positively identified being Arg-Pro.  相似文献   

13.
Monoclonal antibody 13 alpha C5-1-A11 immunoprecipitated two major polypeptides of molecular weights 108,000 and 120,000 from extracts of herpes simplex virus type 2-infected BHK-21 cells labeled with [35S]methionine or [3H]glucosamine. In pulse-chase experiments, both labels were chased from the 120,000-molecular-weight peptide (120K peptide) into the 108K molecule. Endoglycosidase H (endo H) reduced the 120K peptide to a 112K peptide but did not affect the 108K peptide. Similar profiles were obtained with monoclonal antibody AP-1 which reacts with a 92K glycoprotein, gG, which maps to the short unique region of the genome. Cross-absorption experiments indicated that both antibodies reacted with the same peptides, suggesting that the 120K peptide is a partially glycosylated high-mannose-type precursor of gG (pgG1). Immunoprecipitation from monensin-treated cells indicated that pgG1(120K) may undergo peptide cleavage to form a 74K high-mannose-type peptide (pgG2) and that this 74K peptide may be further processed into an endo H-resistant 110K to 116K peptide. In the presence of tunicamycin, gG(108K) was replaced by 110K and 105K peptides which were resistant to both endo H and endoglycosidase F. The 105K peptide was the only molecule labeled by [3H]galactose or [3H]glucosamine in the presence of tunicamycin, and none of the peptides were labeled with [3H]mannose, indicating the probable presence of O-linked sugars in the 105K peptide. Our results imply that cotranslational glycosylation of the unglycosylated precursor 110K peptide results in the high-mannose-type pgG1(120K), which probably undergoes peptide cleavage. This putative cleavage product may then mature into gG (108K) by the trimming of sugars and the addition of complex and probably O-linked sugars; the high-mannose-type pgG2(74K) is probably an intermediate peptide formed in this process.  相似文献   

14.
Acrolein, a known mutagen, undergoes reaction in vitro under physiological conditions with both 2(')-deoxyguanosine and native DNA to give rise to exocyclic adducts of the 5,6,7,8-tetrahydropyrimido[1,2-a]purine-10(3H)-one class having an hydroxy group at either the 6 or the 8 position. Previously we have shown that the 8-hydroxy derivative in a bacterial system has very low mutagenicity probably because in double-stranded DNA this residue exists in the open-chain aldehydic form [N(2)-(3-oxopropyl)-2(')-deoxyguanosine] (3). To continue our investigation in this area, we needed ample supplies of the 6-hydroxy isomers. This current paper describes high-yield simple methods for the synthesis in bulk of the 6-hydroxy adduct 1 and its incorporation into DNA oligomers. The basic methods for the synthesis of the adduct 1, involve 1-substitution of dG derivatives with a 3-butenyl group, dihydroxylation of the olefin with osmium tetroxide and N-methylmorpholine N-oxide, then diol cleavage with periodate ion after incorporation of the 1-(3,4-diacetoxybutyl)-2(')-deoxyguanosine into oligomeric DNA.  相似文献   

15.
Extracts of Klebsiella pneumoniae convert 5-S-methyl-5-thio-D-ribose (methylthioribose) to methionine and formate. To probe the terminal steps of this biotransformation, [1-13C]methylthioribose has been synthesized and its metabolism examined. When supplemented with Mg2+, ATP, L-glutamine, and dioxygen, cell-free extracts of K. pneumoniae converted 50% of the [1-13C]methylthioribose to [13C]formate. The formation of [13C]formate was established by 13C and 1H NMR spectroscopy studies of the purified formate, and by 13C and 1H NMR spectroscopy and mass spectrometry studies of its p-phenylphenacyl derivative. By contrast, no incorporation of label from [1-13C]methylthioribose into the biosynthesized methionine was detected by either mass spectrometry or 13C and 1H NMR spectroscopy. The most reasonable interpretation of these results is that C-1 of methylthioribose is converted directly to formate concomitant with the conversion of carbon atoms 2-5 to methionine. The penultimate step in the conversion of methylthioribose to methionine and formate is an oxidative carbon-carbon bond cleavage reaction in which an equivalent of dioxygen is consumed. To investigate the fate of the dioxygen utilized in this reaction, the metabolism of [1-13C]methylthioribose in the presence of 18O2 was also examined. Mass spectrometry revealed the biosynthesis of substantial amounts of both [18O1]methionine and [13C, 18O1]formate under these conditions. These results suggest that the oxidative transformation in the conversion of methylthioribose to methionine and formate may be catalyzed by a novel intramolecular dioxygenase. A mechanism for this dioxygenase is proposed.  相似文献   

16.
Aqueous solutions of adenosylcobalamin (AdoCbl) were thermolyzed with excess beta-mercaptoethanol under anaerobic conditions. The product studies reveal that approximately 90% Co-C bond homolysis occurs, to yield Co(II)cobalamin, 5'-deoxyadenosine, and the disulfide product from the combination of two HOCH2CH2S* radicals, 2,2'-dithiodiethanol; there is also approximately 10% Co-C bond heterolysis, yielding Co(III)cobalamin, adenine, and 2,3-dihydroxy-4-pentenal. The kinetic studies show there is a first-order dependence on AdoCbl and zero-order dependence on thiol under the higher [RSH] experimental conditions used, consistent with the rate-determining step at high [RSH] being the generation of Ado*. The kinetic results require that, in enzyme-free AdoCbl solution, adenosyl radical (Ado*) is formed as a discrete intermediate which then abstracts H* from the added thiol. The activation parameters for Co-C bond homolysis in the presence of thiol trap are the same within experimental error as the activation parameters for Co-C bond homolysis without trap, standard delta H(obs) = 29(2) kcal mol(-1) and standard delta S(obs) = -1(5) e.u. The results, in comparison to the rate of Co-C bond homolysis in ribonucleoside triphosphate reductase (RTPR), reveal that RTPR accelerates Co-C bond cleavage in AdoCbl by approximately 10(10+/-1). The recent literature evidence bearing on the exact mechanism of RTPR enzymic cleavage of the Co-C bond of AdoCbl is briefly discussed, notably the fact that this mechanism is presently controversial, but does involve at least coupled (and possibly concerted) Co-C cleavage, -S-H cleavage, and C-H (Ado-H) formation steps.  相似文献   

17.
Following incubation of [3H]dynorphin A (1-8) and [3H]dynorphin A (1-9) with suspensions of guinea pig brain membranes, analysis of the supernatants by HPLC has shown that both peptides are degraded at 25 degrees C and at 0 degrees C. Bestatin and captopril reduce degradation at 0 degrees C but for a similar degree of protection at 25 degrees C arginine-containing dipeptides are also required. The effects of these peptidase inhibitors on the degradation profiles indicate that [3H]dynorphin A (1-8) has three main sites of cleavage: the Tyr1-Gly2, Arg6-Arg7, and Leu5-Arg6 bonds. With [3H]dynorphin A (1-9) as substrate the Arg7-Ile8 and Ile8-Arg9 bonds are also liable to cleavage. In binding assays, in contrast to the effects of peptidase inhibitors on the degradation of unbound [3H]dynorphin A (1-8) and [3H]dynorphin A (1-9), bestatin and captopril have little effect on the binding characteristics of the tritiated dynorphin A fragments at the kappa-site at 0 degrees C. However, at 25 degrees C binding is low in the absence of peptidase inhibitors. When binding at mu- and delta-sites is prevented, the maximal binding capacities of [3H]dynorphin A (1-8), [3H]dynorphin A (1-9), and [3H](-)-bremazocine at the kappa-site are similar; [3H]dynorphin A (1-9) has 5-10 times higher affinity for the kappa-site than [3H]dynorphin A (1-8). Comparison of the effects of peptidase inhibitors on unbound dynorphin A fragments with their effects in binding assays suggests that the bound peptides are protected from the action of peptidases.  相似文献   

18.
M F Farnum  J P Klinman 《Biochemistry》1986,25(20):6028-6036
Bovine plasma amine oxidase (PAO) has previously been shown to catalyze a nonstereospecific loss of tritium from [2(R)-3H]- and [2(S)-3H]dopamines, attributed to multiple, catalytically active binding sites for substrate [Summers, M. C., Markovic, R., & Klinman, J. P. (1979) Biochemistry 18, 1969-1979]. Analysis of products formed from incubation of dopamine with PAO in tritiated water indicates a stereospecific, pro-R, incorporation of label at C-2. Thus, tritium washout (random) and washin (pro-R) are not the microscopic reverse of one another. We conclude that the (enamine) intermediates leading to tritium washin are nonequivalently bound. The observation of pro-R incorporation has provided a straightforward synthetic route to [1(R)-2H,2(R)-3H]- and [1(S)-2H,2(R)-3H]dopamines, which upon oxidation with PAO are expected to be processed preferentially by 1S and 1R cleavage, respectively. From previously measured isotope effects, we predict the loss of tritium from the 1(R)-2H and 1(S)-2H samples to be 74:8 for a syn relationship between cleavage at C-1 and C-2 vs. 21:90 for an anti relationship. The observation of a 68:18 ratio at 100% conversion provides strong evidence for a syn cleavage. The data support a mechanism in which a single base catalyzes a 1,3-prototrophic shift of hydrogen from C-1 of the substrate to cofactor, followed by exchange from C-2. Additionally, the results confirm the presence of alternate binding modes for dopamine at the active site of bovine plasma amine oxidase. This interaction of dopamine with plasma amine oxidase is a rare example of mirror-image catalysis in which a single substrate has two functional binding orientations on an enzyme surface.  相似文献   

19.
Allylic addition-elimination reactions are widely used in the enzyme-catalysed formation of terpenoid metabolites. It has earlier been shown that the isoprenoid chain elongation reaction catalysed by farnesyl pyrophosphate synthase involving successive condensations of dimethylallyl pyrophosphate (DMAPP) and geranyl pyrophosphate (GPP) with isopentenyl pyrophosphate (IPP) corresponds to such an SE' reaction with net syn stereochemistry for the sequential electrophilic addition and proton elimination steps. Studies of the enzymic cyclization of farnesyl pyrophosphate (FPP) to pentalenene have now established the stereochemical course of two additional biological SE' reactions. Incubation of both (9R)- and (9S)-[9-3H,4,8-14]FPP with pentalenene synthase and analysis of the resulting labelled pentalenene has revealed that H-9re of FPP becomes H-8 of pentalenene, while H-9si undergoes net intramolecular transfer to the adjacent carbon, becoming H-1re (H-1 alpha) of pentalenene, as confirmed by subsequent experiments with [10-2H, 11-13C]FPP. These results correspond to net anti-stereochemistry in the intramolecular allylic addition-elimination reaction. The stereochemical course of a second SE' reaction has now been examined by analogous incubations of (4S,8S)-[4,8-3H,4,8-14C]FPP and (4R,8R)-[4,8-3H, 4.8-14C]FPP with pentalenene synthase. Determination of the distribution of label in the derived pentalenenes showed stereospecific loss of the original H-8si proton. Analysis of the plausible conformation of the presumed reaction intermediates revealed that the stereochemical course of the latter reaction cannot properly be described as either syn or anti, since cyclization and subsequent double bond formation require significant internal motions to allow proper overlap of the scissile C-H bond with the developing carbocation.  相似文献   

20.
the synthesis and in vivo application of stable-isotopically labeled folic acid was investigated to devise methods suitable for studies of folate metabolism in human subjects. Glutamate-labeled tetradeutero-pteroylglutamic acid (d4-folic acid) was prepared by mixed anhydride coupling of N10-trifluoroacetylpteroic acid and dimethyl L-[3,3,4,4-2H4]glutamic acid, saponification in sodium deuteroxide, and chromatographic purification. Retention of the isotopic label was verified by proton NMR and mass spectrometry of the para-aminobenzoylglutamic acid product of C9-N10 bond cleavage. A method was devised for determination of of isotopic enrichment of urinary d4-folates derived from orally administered d4-folic acid using affinity chromatographic purification, chemical cleavage of the C9-N10 bond, HPLC isolation of the p-[2H4]aminobenzoylglutamate product, followed by negative-ion chemical-ionization gas chromatography/mass spectrometry. Data concerning the urinary excretion of d4-folates derived from an oral dose of d4-folic acid in an adult human are presented.  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