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.
These studies indicate that the interconversions of delta 1-pyrroline-5-carboxylate and proline can function as a shuttle that generates extra-mitochondrial NADP+ and transfers hydride ions into mitochondria in a cell-free rat liver system. A phosphate-free buffer with high concentrations of triethanolamine and 2-mercaptoethanol prevented the cold inactivation of pyrroline-5-carboxylate reductase (EC 1.5.1.2) in liver extracts. This enzyme had an apparent KmNADPH that was 2% of the apparent KmNADH X VmaxNADPH was approx. 50% of VmaxNADH. Unlabeled proline was converted to [5-3H]proline in incubations containing liver soluble fraction, mitochondria and a [4S-3H]NADPH generating system. This demonstrated one turn of the proposed shuttle in a homologous liver system. [5-3H]Proline production increased linearly over 60 min and decreased by 87% or more when specific components were eliminated. Rotenone was required for maximal activity, suggesting that inhibition of delta 1-pyrroline-5-carboxylate efflux would be required for significant shuttle activity in vivo. Both the relative concentrations of NADPH and NADH in liver cytosol and the kinetic characteristics of liver pyrroline-5-carboxylate reductase predict that the described shuttle should be overwhelmingly linked to NADPH rather than NADH. A NADPH-linked delta 1-pyrroline-5-carboxylate-proline shuttle may occur in hepatocytes and function at specific times to regulate pathways limited by cytosolic [NADP+].  相似文献   

2.
Direct evidence is presented for a proline cycle using a cell-free experimental system which sequentially transfers 3H from [1-3H]glucose to NADP+ to Δ1-pyrroline-5-carboxylate and yields [3H]proline. The formation of [3H]proline depends on the presence of NADP, Δ1-pyrroline-5-carboxylate, and the enzymes glucose-6-phosphate dehydrogenase and Δ1-pyrroline-5-carboxylate reductase. The production of [3H]proline from unlabeled proline in the presence of mitochondria provides direct evidence for one complete turn of a proline cycle which transfers reducing equivalents produced by glucose oxidation in the pentose pathway into mitochondria. In this cycle, proline is oxidized to Δ1-pyrroline-5-carboxylate by mitochondrial proline oxidase. Δ1-pyrroline-5-carboxylate is released from mitochondria and is recycled back to proline by Δ1-pyrroline-5-carboxylate reductase with concomitant oxidation of NADPH. At the maximal rate observed, 60% of Δ1-pyrroline-5-carboxylate produced is recycled back to proline. This cycle provides a mechanism for transferring reducing equivalents from NADPH into mitochondria and is linked to glucose oxidation in the pentose pathway by NADPH turnover.  相似文献   

3.
R J Smith 《Enzyme》1984,31(2):115-121
A radioisotopic assay is described for measuring the activity of delta 1-pyrroline-5-carboxylate synthase, the enzyme that catalyzes the formation of delta 1-pyrroline-5-carboxylic acid from glutamic acid. Pyrroline-5-carboxylic acid is a common intermediate in the pathways through which glutamic acid, proline, and ornithine are interconverted. To determine pyrroline-5-carboxylate synthase activity, cell homogenates are incubated with [14C]-glutamic acid, the products of the reaction are converted quantitatively to proline by sodium borohydride, and proline is isolated by cation-exchange column chromatography. Cofactor requirements have been defined, and the activity of pyrroline-5-carboxylate synthase in several different cultured fibroblast lines is reported.  相似文献   

4.
A radiochemical assay has been developed for a NADP+-specific gamma-glutamate semialdehyde dehydrogenase from rat intestinal epithelial cells. The spectrophotometric assay utilized to measure the enzyme in bacterial cell homogenates is not sensitive enough for homogenates from rat mitochondria, which require an assay that can measure as little as 0.5 nmol NADPH formed/min/ml extract. The assay described here is sensitive to 0.1 nmol product formed/min/ml of extract and employs the use of [3H]pyrroline 5-carboxylate which is phosphorylated and oxidized by the enzyme to gamma-[3H]glutamyl phosphate, a product that decomposes to [3H]pyrrolidone 5-carboxylate. The latter product is separated from the substrate by ion-exchange chromatography. In order to correct for any product loss during separation by ion-exchange [14C]pyrrolidone 5-carboxylate is added as an internal standard to the deproteinized assay mixture. Under the assay conditions described mammalian gamma-glutamate semialdehyde dehydrogenase activity is linear with respect to time and protein concentration. Comparison between the kinetic parameters reported for the bacterial enzyme and those reported here for the mammalian enzyme indicate similarities in the pH optima as well as a requirement for phosphate. Kinetic studies on mammalian enzyme yield apparent Km values of 1.8 mM for pyrroline 5-carboxylate, 0.2 mM for NADP+, and 11.3 mM for phosphate.  相似文献   

5.
A rapid and simple method for assay of pyrroline-5-carboxylate synthase is presented. In this method, the incubation is terminated by raising the pH of incubation mixture to 10, and [14C]pyrroline 5-carboxylate produced from the substrate, [14C]glutamate, is first converted quantitatively to [14C]proline by reduction with NaBH4 at pH 10 and then the proline is allowed to pass through column of AG1-X8 anion exchanger under the conditions where the glutamate is completely retained by the column. Radioactive counting of the eluate gives the synthase activity. The entire procedure takes only one hour.  相似文献   

6.
The objective of these experiments was to determine the fate of tritium from the 5 position of proline and to assess the validity of its loss to H2O as a measure of proline oxidation. When [5-3H]proline was fed to barley (Hordeum vulgare) leaves, tritium was recovered in H2O and metabolites such as glutamate, glutamine, organic acids, aspartate, asparagine, and γ-aminobutyrate. Collectively these metabolites, which are oxidation products of proline, accounted for 8% of the 3H recovered after 5 hours. In spite of the amount recovered in metabolites, the rates of proline oxidation estimated by measuring 3H2O recovery from [5-3H]proline were only slightly lower than rates estimated by incorporation of 14C into oxidized products and loss of 14C from total proline. Therefore, 3H2O recovery from [5-3H]proline is useful in assessing the effects of stress on proline metabolism.

Water stress inhibited proline oxidation, as reported previously. In addition, a reconversion of proline oxidation products to proline occurred in stressed leaves. This observation probably indicates a breakdown in cellular compartmentation of proline synthesis and proline oxidation.

  相似文献   

7.
We describe a radioisotopic assay for Δ1-pyrroline-5-carboxylate reductase. In this assay we use Δ1-pyrroline-5-carboxylate[U-14C] and isolate product l-[U-14C]proline by cation-exchange column chromatography.  相似文献   

8.
[1'-3H]- and [2'-3H]dihydroneopterin triphosphate (NH2TP) were prepared enzymatically from [4-3H]- and [5-3H]glucose and converted to tetrahydrobiopterin (BH4) by an extract from bovine adrenal medulla. The formation of BH4 from both [1'-3H]- and [2'-3H]-NH2TP proceeds with virtually complete loss of the respective tritium label. The breaking of the CH-bond at C-1' is characterized by a kinetic isotope effect of 2.6 +/- 0.5. A smaller kinetic isotope effect of 1.5 +/- 0.2 was found for the breaking of the CH-bond at C-2'.  相似文献   

9.
In the course of studies on the oxygenation of steroids by purified P-450 cytochromes, particularly rabbit liver microsomal cytochrome P-450 form 3b, a rapid and reliable radiometric assay has been devised for progesterone 16 alpha-hydroxylation. In view of the lack of a commercially available, suitably tritiated substrate, [1,2,6,7,16,17-3H]progesterone was treated with alkali to remove the label from potential hydroxylation sites other than the 16 alpha position. The resulting [1,7,16-3H]progesterone was added to a reconstituted enzyme system containing cytochrome P-450 form 3b, NADPH-cytochrome P-450 reductase, and NADPH, and the rate of 16 alpha-hydroxylation was measured by the formation of 3H2O. This reaction was shown to be linear with respect to time and to the cytochrome P-450 concentration. An apparent tritium isotope effect of 2.1 was observed by comparison of the rates of formation of tritium oxide and 16 alpha-hydroxyprogesterone, and the magnitude of the isotope effect was confirmed by an isotope competition assay in which a mixture of [1,7,16-3H]progesterone and [4-14C]progesterone was employed.  相似文献   

10.
Fatty acids containing a prochiral tritium label have often been used in the study of enzymatic reactions which involve an obligatory step of hydrogen abstraction. In the lipoxygenase reaction, the primary isotope effect associated with this approach is detected as an isotopic enrichment of the substrate. Herein we characterize a previously unrecognized secondary isotope effect which changes the specific activity of both the substrate and product. The 12-lipoxygenase of human platelets removes the 10-LS hydrogen of arachidonic acid in the formation of 12-hydroperoxyeicosatetraenoic acid. We studied the specific activity changes associated with conversion of the enantiomerically labeled [10-DR-3H]arachidonic acid to 12-[10-3H]hydroxyeicosatetraenoic acid in aspirin-treated platelets. [3-14C]Arachidonic acid served as internal standard. The most pronounced change in 3H/14C ratio in the early stages of reaction was a 15-20% deficiency of tritium in the product. Later, the remaining arachidonate showed a marked increase in 3H/14C ratio. The changes in specific activity closely matched those predicted for a secondary isotope effect. Comparison of these data with the theoretical equations for a secondary isotope effect indicated the 10-DR-3H substrate reacted at about 84% of the rate of unlabeled molecules. Interestingly, this secondary isotope effect is similar in magnitude to the secondary isotope effect in autoxidation reactions, a finding compatible with a basic similarity in reaction mechanisms in enzymatic and non-enzymatic oxygenation of lipids.  相似文献   

11.
The biosynthesis of S-adenosylmethionine occurs in a unique enzymatic reaction in which the synthesis of the sulfonium center results from displacement of the entire polyphosphate chain from MgATP. The mechanism of S-adenosylmethionine synthetase (ATP:L-methionine s-adenosyltransferase) from Escherichia coli has been characterized by kinetic isotope effect and substrate trapping measurements. Replacement of 12C by 14C at the 5' carbon of ATP yields a primary Vmax/Km isotope effect (12C/14C) of 1.128 +/- 0.003 in the absence of added monovalent cation activator (K+). At saturating K+ concentrations (10 mM) the primary isotope effect diminishes slightly to 1.108 +/- 0.003, indicating that the step in the mechanism involving bond breaking at the 5' carbon of MgATP has a small commitment to catalysis at conditions near Vmax. No alpha-secondary 3H isotope effect from [5'-3H]ATP was detected, (1H/3H) = 1.000 +/- 0.002, even in the absence of KCl. There was no significant primary sulfur isotope effect from [35S]methionine at KCl concentrations from 0 to 10 mM. Substitution of the methyl group of methionine with tritium yielded a beta-secondary isotope effect (CH3/C3H3) = 1.009 +/- 0.008 independent of KCl concentration. The reaction of selenomethionine and [5'-14C]ATP gave a primary isotope effect of 1.097 +/- 0.006, independent of KCl concentration. Substrate trapping experiments demonstrated that the step in the mechanism involving bond making to sulfur of methionine does not have a significant commitment to catalysis at 0.25 mM KCl, therefore intrinsic isotope effects were observed. Substrate trapping experiments indicated that the step involving bond breaking at carbon 5' of MgATP has a 10% commitment to catalysis at 0.25 mM KCl. The isotope effects are interpreted in terms of an Sn2-like transition state structure in which bonding of the C5' is symmetric with respect to the departing tripolyphosphate group and the incoming sulfur of methionine. With selenomethionine as substrate an earlier transition state is implicated.  相似文献   

12.
Zhu Y  Shearer G  Kohl DH 《Plant physiology》1992,98(3):1020-1028
Supplying l-proline to the root system of intact soybean (Glycine max [L.] Merr.) plants stimulated acetylene reducing activity to the same extent as did supplying succinate. Feeding l-proline also caused an increase in bacteroid proline dehydrogenase activity that was highly correlated with the increase in acetylene-reducing activity. Twenty-four hours after irrigating with l-proline, endogenous proline content had increased in host cell cytoplasm and bacteroids, about three- and eightfold, respectively. In bacteroids, proline concentration was calculated to be at least 3.5 millimolar. In experiments in which [U-14C]l-proline was supplied to uprooted, intact plants incubated in aerated solution, 14C-labeled products of proline metabolism, as well as [14C]proline itself, accumulated in both host cells and bacteroids. When plants were incubated in aerated solutions containing [5-3H]l-proline, 3H-labeled proline was found in host cells and bacteroids. [3H] Pyrroline-5-carboxylate was found in bacteroids, but not host cells, after a 2-hour incubation in [5-3H]l-proline. When [U-14C]l-proline was supplied for 24 hours, a significant amount of [14C] pyrroline-5-carboxylate was found in the host cells, in contrast with the results from the shorter incubation in [5-3H]proline, although the amount in the host cells was only about half the quantity found in the bacteroids. Taken as a whole, these results indicate that proline crosses both plant and bacterial membranes under the in vivo experimental conditions utilized and are consistent with a significant role for proline as an energy source in support of bacteroid functioning. In spite of the increase in acetylene-reducing activity when proline was supplied to the root system of intact plants, proline application did not rescue stemgirdled plants from loss of acetylene-reducing activity, although succinate application did. This suggests a nonphloem route for succinate, but not proline, from roots to nodules.  相似文献   

13.
The stereochemical course of the aliphatic hydroxylation of gamma-butyrobetaine by calf liver and by Pseudomonas sp AK1 gamma-butyrobetaine hydroxylases has been determined. With [3(RS)-3-3H]-gamma-butyrobetaine or [3(R)-3-3H]-gamma-butyrobetaine as substrate, a rapid and significant loss of tritium to the medium occurred. On the other hand, with [3(S)-3-3H]-gamma-butyrobetaine, only a negligible release of tritium to the aqueous medium was observed. Indeed, on hydroxylation of [3(S)-3-2H]-gamma-butyrobetaine by either the calf liver or bacterial hydroxylase, the isolated product L-carnitine was found to have retained all of the deuterium initially present in the 3(S) position. Since the absolute configuration of the product L-carnitine has been determined to be R, such results are only compatible with a hydroxylation reaction that proceeded with retention of configuration. With [methyl-14C,3(R)-3-3H]-gamma-butyrobetaine as substrate for the calf liver hydroxylase, the percentage of tritium retained in the [methyl-14C]-L-carnitine product was determined as a function of percent reaction. The results of these studies indicated that pro-R hydrogen atom abstraction exceeded 99.9%. Experiments using racemic [methyl-14C,3(RS)-3-3H]-gamma-butyrobetaine as substrate yielded similar results and additionally allowed us to estimate alpha-secondary tritium kinetic isotope effects of 1.10 and 1.31 for the bacterial and calf liver enzymes, respectively. These results are discussed within the context of the radical mechanism for gamma-butyrobetaine hydroxylase previously proposed [Blanchard, J. S., & Englard, S. (1983) Biochemistry 22, 5922], and the required topographical arrangement of enzymic oxidant and substrate is illustrated.  相似文献   

14.
The stereochemistry of the bovine plasma amine oxidase catalyzed oxidation of 2-(3,4-dihydroxyphenyl)-ethylamine (domapine) has been investigated by comparing 3H/14C ratios of 3,4-dibenzyloxyphenethyl alcohols, derived from 3,4-dihydroxyphenylacetaldehydes, to starting dopamines chirally labeled at C-1 and C-2. The oxidation of [2RS-3H]-, [2R-3H]-, and [2S-3H]dopamine leads to products which have retained 53, 59, and 47% of their tritium. Similarly, oxidation of [1RS-3H]-, [1R-3H]-, and [1S-3H]dopamine leads to an 80, 80, and 92% retention of tritium. The configurational purity of tritium at C-2 of dopamine and C-1 of the dopamine precursor 3-methoxy-4-hydroxyphenethylamine has been confirmed employing dopamine-beta-hydroxylase (specific for the pro-R hydrogen at C-2) and pea seedling amine oxidase (specific for the pro-S hydrogen at C-1). In addition, chromatographically resolved isozymes of bovine plasma amine oxidase have been demonstrated to lead to the same stereochemical result as pooled enzyme fractions. We have been able to rule out carbon interchange and tritium transfer in the ethylamine side chain of dopamine as the source of the apparent nonstereospecificity. Estimated primary tritium isotope effects are 1 for [2-3H]dopamines and 5--6 and 26--34 for [1R-3H]- and [1S-3H]dopamine, respectively. We propose the presence of alternate dopamine binding modes, characterized by absolute but opposing stereochemistries and differential primary tritium isotope effects at C-1.  相似文献   

15.
The rate of tritium removal from l[3-3H]lactate by hamster liver cells is faster than the analytical rate of lactate utilization, or the rate of 14C disappearance from l[U-14C]lactate, with the result that the 3H/14C ratio in residual lactate from l-[U-14C,3-3H]lactate decreases. However, addition of low concentrations (0.1 to 1.0 mM) of l-cycloserine, a glutamate pyruvate transaminase inhibitor, nearly equalizes the rates of isotope utilization from l-[3-3H]lactate and l-[U-14C]lactate. The results suggest a very limited rate of recycling of phosphoenolpyruvate back to pyruvate during gluconeogenesis from lactate in fasted hamster liver cells.  相似文献   

16.
1. (2R,3S)-[U-14C,3-3H1]- and (2R,3R)-[U-14C,2,3-3H2] Cysteine hydrochlorides have been separately synthesised. The latter compound has been shown to have uniform distributions of tritium between C-2 and C-3. 2. The abvoe cysteines and (2R)-[U-14C,3,3,3',3'-3H4]cystine have been converted to samples of penicillin G by Penicillium chrysogenum. 3. Incorporation results indicate that all but 14% of the tritium is lost from the (2R,3S)-[3-3H1]isomer; that 42% of tritium is retained by the non-stereospecifically C-3 tritiated cystine; and that 58% of tritium is retained by the (2R,3R)-[2,3-3H2]isomer on conversion to penicillin G. 4. Degradation of the penicillin G derived from (2R,3R)-[U-14C,2,3-3H2]cysteine hydrochloride has indicated that in fact about 87% of the original C-3 tritium of cysteine is retained at C-5 of penicillin G. 5. The results indicate stereospecificity in the cyclisation giving rise to the beta-lactam ring in penicillin G in nature with loss of the 3-pro-S-hydrogen and rentention of the 3-pro-R-hydrogen of cysteine. Thus there is net retention of stereochemistry in the cyclisation.  相似文献   

17.
The multiple isotope effect method of Hermes et al. [Hermes, J. D., Roeske, C. A., O'Leary, M. H., & Cleland, W. W. (1982) Biochemistry 21, 5106-5114] has been used to study the mechanism of the oxidative decarboxylation catalyzed by 6-phosphogluconate dehydrogenase from yeast. 13C kinetic isotope effects of 1.0096 and 1.0081 with unlabeled or 3-deuterated 6-phosphogluconate, plus a 13C equilibrium isotope effect of 0.996 and a deuterium isotope effect on V/K of 1.54, show that the chemical reaction after the substrates have bound is stepwise, with hydride transfer preceding decarboxylation. The kinetic mechanism of substrate addition is random at pH 8, since the deuterium isotope effect is the same when either NADP or 6-phosphogluconate or 6-phosphogluconate-3-d is varied at fixed saturating levels of the other substrate. Deuterium isotope effects on V and V/K decrease toward unity at high pH at the same time that V and V/K are decreasing, suggesting that proton removal from the 3-hydroxyl may precede dehydrogenation. Comparison of the tritium effect of 2.05 with the other measured isotope effects gives limits of 3-4 on the intrinsic deuterium and of 1.01-1.05 for the intrinsic 13C isotope effect for C-C bond breakage in the forward direction and suggests that reverse hydride transfer is 1-4 times faster than decarboxylation.  相似文献   

18.
The isotope effects for the interconversion of L-proline and D-proline, catalyzed by proline racemase, have been determined in the saturated region with both [2-2H]proline and [2-3H]proline. The deuterium fractionation factors for each of the protons in flight have been obtained from two kinds of experiment: by measuring the rate of racemization of one [2-2H]proline enantiomer as it racemizes into an equilibrated pool of unlabeled proline and by measuring the deuterium content of a proline sample at the optical rotation maximum that occurs when an equimolar mixture of one deuterium-labeled enantiomer and the other unlabeled enantiomer runs to equilibrium. The tritium fractionation factors for each of the protons in flight have been determined from measurements of the rate of loss of tritium to the solvent as one [2-3H]proline enantiomer runs to equilibrium. Good agreement is found among the fractionation factors determined by each method. The deuterium fractionation factors for the two protons are not identical: that for the proton derived from L-proline is 0.375 and that for the proton derived from D-proline is 0.44. This difference has been confirmed by a double-competition experiment in which the optical rotation of a mixture of DL-[2-2H]proline and unlabeled DL-proline is followed with time. The rotation (initially zero) passes through a maximum, from which the ratio of the two fractionation factors (0.86) is obtained. These data, coupled with the equilibrium fractionation factor for the 2-position of proline (which has been determined to be 1.17), provide the transition-state factors for each of the in-flight protons, and delineate the nature of the transition state(s) for the enzyme-catalyzed racemization.  相似文献   

19.
1. (3RS,6R)-[6-2H1,6-3H1,6-14C], (3RS,6S)-[6-2H1,6-3H1,6-14C] and (3RS)-[6-3H1,6-14C]mevalonolactones were synthesised from R-[2H1,3H1,2-14C], S-[2H1,3H1,2-14C] and [3h1,2-14C]acetic acids respectively. 2. Each mevalonate was converted into cholesterol by a rat liver preparation. 3. Each cholesterol specimen was converted into androsta-1,4-diene-3,17-dione by incubation with Mycobacterium phlei in the presence of 2,2'.dipyridyl. Each specimen of androsta-1,4-diene-3,17-dione was converted into androsta-1,4-dien-3-one-17-ethylene ketail. 4. The samples of androsta-1,4-dien-3-one-17-ethylene ketal were each converted chemically into oestrones in which the methyl group at C-18 is the only carbon atom that originated from C-6 in mevalonolactone. 5. The oestrone from (3RS)-[6-3H1,6-14C]mevalonolactone was oxidised chemically to acetic acid which was converted into p-bromophenacyl acetate and the 3H/14C ratio was measured. 6. There was no overall loss of tritium from the methyl group of acetic acid, as measured by determining the 3H/14C ratios of the p-bromophenacyl esters, when the synthetic and degradative procedures 1 -- 5 were tested with [3H1,2-14C]acetic acid. 7. The oestrones derived from the 6R and 6S-mevalonolactones were oxidised. The chiralities of the resulting acetates were determined by an established procedure whereby the acetates were converted into 2S-malates which were examined for loss of tritium on equilibration with fumarate hydratase. 8. The oestrone from (3RS,6R)-[6-2H1,6-3H1,6-14C]mevalonate gave acetic acid which was converted into 2S-malate that retained 68.6% of its tritium after treatment with fumarate hydratase; the configuration of this acetic acid was R. 9. The oestrone from (3RS,6S)-E16-2H1,6-3H1,6-14C]mevalonate was oxidised to acetic acid which was converted into 2S-malate that retained 31.9% of its tritium after treatment with fumarate hydratase; the configuration of this acetic acid was S. 10. There was no overall change in the configuration of a chiral methyl group between C-6 of mevalonate and C-18 of oestrone. It is cncluded that the intramolecular migration of a chiral methyl group from C-15 in 2,3-oxidosqualene to C-13 in lanosterol is stereospecific and occurs with overall retention of configuration.  相似文献   

20.
In order to study the initial as well as the final steps in the aromatization of androgens to estrogens, high-specific activity [19-C3H3]androstenedione and testosterone were synthesized. Incubations of [19-C3H3]androstenedione with human placental microsomes resulted in the generation of [3H]water, as a result of the dual hydroxylation at C-19, and [3H]formic acid reflecting final aromatization. After an initial lag in the production of [3H]formic acid, the two radiolabeled products were formed linearly with time at a ratio of 2 to 1 under subsaturating conditions and 2.2 to 1 when saturating levels of substrate were present. Incubation of a mixture of [19-C3H3]- and [4-14C]androstenedione with human placental microsomes yielded 19-hydroxy- and 19-oxoandrostenedione, respectively, products of one and two hydroxylations at C-19. The isotope ratios of these derivatives revealed the presence of a tritium isotope effect in the first but not in the second hydroxylation at that site. When [19-C3H2]- and [4-14C]19-hydroxyandrostenedione were used as the substrate, the isotope ratio of the isolated 19-oxoandrostenedione showed no evidence of any isotope effect in its formation. Thus, the second hydroxylation at C-19 exhibits no isotope effect irrespective of whether androstenedione or 19-hydroxyandrostenedione are the substrates, and therefore, a concerted process and catalytic commitment are not responsible for the difference in isotope effects between the first and second C-19 hydroxylation by the placental aromatase complex. Radiometric kinetic analysis employing [19-C3H3]- and [1 beta,2 beta-3H]androstenedione as the comparative substrates provided evidence that the isotope effect is exerted solely through the Vmax component of the reaction. The distinction between the successive hydroxylations at C-19 in the aromatization sequence suggests, but does not prove, that different mechanisms, and hence different catalytic sites, may be involved in these steps.  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