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 46 毫秒
1.
Three factors are of primary importance with respect to designing efficient P450 biocatalysts. (1) The substrate must be oxidized at a significant rate. (2) The regioselectivity must heavily favor the desired product. (3) The enzyme must use the majority of the reducing equivalents from NADH or NADPH to produce product. The reaction we chose to study was oxidation of 2-ethylhexanol to 2-ethylhexanoic acid by P450cam. We examined four active site mutations: F87W, Y96W, T185F, and L244A. The mutations were chosen to improve 2-ethyhexanoic acid production by decreasing active site volume, increasing active site hydrophobicity, and improving stereoselectivity. The F87W and Y96W mutations improved regioselectivity, giving almost exclusively the desired product. The T185F mutation improved coupling of NADH to product formation. The L244A mutation altered the stereoselectivity of 2-ethylhexanoic acid production. These results indicate that active site mutations of P450cam can alter catalysis of 2-ethylhexanol.  相似文献   

2.
The stereoselectivity of the peroxisome proliferation potency of 2-ethylhexanoic acid (2-EHA), a metabolite of the plasticizer di-(2-ethylhexyl) adipate, was investigated in vitro. The enantiomers of 2-EHA were prepared via the semipreparative HPLC resolution of their diastereoisomeric (+)-(R)-1-phenylethylamine derivatives and the subsequent hydrolytic cleavage. Monolayers of hepatocytes were incubated 3 days with solution of (-)-(R), (+)-(S), and (+/-)-2-EHA. The peroxisome proliferation potency was measured by means of determination of the peroxisomal palmitoyl coenzyme A oxidation. The theoretical induction component due to each enantiomer were calculated from the experimental data considering the enantiomeric purities of the acids. The (+)-(S)-enantiomer was found to be the most potent inducer e.g., the eutomer, while the (-)-(R) was the distomer. The eudismic ratio was about 1.6 and the racemic mixture exhibited an intermediary potency. These results, obtained in vitro in conditions avoiding confounding factors such as pharmacokinetics, suggest that the peroxisome proliferation induced by 2-ethylhexanoic acid is a stereoselective phenomenon.  相似文献   

3.
Resonance Raman spectroscopy at 2.5cm(-1) resolution was used to probe differences in wild-type and Y96F mutant P450cam (CYP101), both with and without bound camphor or styrene substrates. In the substrate-free state, the spin state equilibrium is shifted from 6-coordinate low spin (6CLS) toward more 5-coordinate high spin (5CHS) when tyrosine-96 in the substrate pocket is replaced by phenylalanine. About 25% of substrate-free Y96F mutant is 5CHS as opposed to 8% for substrate-free wild-type P450cam. Spin equilibrium constants calculated from Raman intensities indicate that the driving force for electron transfer from putidaredoxin, the natural redox partner of P450cam, is significantly smaller on styrene binding than for camphor binding. Spectral differences suggest that there is a tilt in camphor toward the pyrrole III ring on Y96F mutation. This finding is consistent with the altered product distribution found for camphor hydroxylation by the Y96F mutant relative to the single enantiomer produced by the wild-type enzyme.  相似文献   

4.
HA-966 (1-hydroxy-3-aminopyrrolidone-2) is an antagonist at the glycine allosteric site of the N-methyl-D-aspartate receptor ionophore complex. Unlike presently known glycine antagonists, HA-966 is chiral. We report stereoselectivity for the (R)-enantiomer at the glycine antagonist site. In [3H]glycine binding, the (R)-enantiomer has an IC50 of 4.1 +/- 0.6 microM. The racemic mixture has an IC50 of 11.2 +/- 0.5 microM, whereas (S)-HA-966 has an IC50 greater than 900 microM. In glycine-stimulated [3H]1-[1-(2- thienyl)cyclohexyl]piperidine binding, the (R)-enantiomer inhibits with an IC50 of 121 +/- 61 microM, whereas the racemic mixture has an IC50 of 216 +/- 113 microM and (S)-HA-966 is inactive. The inhibition by (R)-HA-966 can be prevented by the addition of glycine. (R)-HA-966 and racemic HA-966, but not (S)-HA-966, also prevent N-methyl-D-aspartate cytotoxicity in cortical cultures. The (R)-enantiomer and, less potently, the (S)-enantiomer inhibit N-methyl-D-aspartate-evoked [3H]norepinephrine release from rat hippocampal slices (IC50 values of about 0.3 mM and 1.6 mM, respectively), but only the inhibition by (R)-HA-966 is reversed by added glycine. In glutamate-evoked contractions of the guinea pig ileum, (R)-HA-966 causes a glycine-reversible inhibition (IC50 of about 150 microM), whereas (S)-HA-966 is much less potent (IC50 of greater than 1 mM). These results demonstrate stereoselectivity of the glycine antagonist site of the N-methyl-D-aspartate receptor complex in a variety of tissues and assays. The stereoselectivity also confirms the specificity of N-methyl-D-aspartate receptors in glutamate-evoked contractions of the guinea pig ileum, and supports their similarity to central N-methyl-D-aspartate receptors.  相似文献   

5.
T L Poulos  R Raag 《FASEB journal》1992,6(2):674-679
Several crystal structures of various substrate and inhibited complexes of the camphor monoxygenase, cytochrome P450cam from Pseudomonas putida, are now available. These structures, together with mutagenesis, biochemical, and biophysical studies, have allowed for a detailed penetration into the problem of how P450s activate molecular oxygen, control stereoselectivity, and transfer electrons. This review will provide a summary of the crystallographic work in light of what these structures have taught us about P450 function.  相似文献   

6.
Cytochrome P450 monooxygenases (P450s) are powerful biocatalysts that have the ability to oxidize a broad range of substrates, often at non-reactive carbon centers. However, incorporation of P450s into synthetic schemes has so far been limited to a few whole-cell transformations. P450 substrates are often hydrophobic and have low water solubility, limiting the amount of product that can be produced. To help overcome this limitation, we have examined P450cam activity in two-phase hexane/water emulsions with and without the anionic surfactant, bis(2-ethylhexyl) sulfosuccinate sodium salt (AOT). Hydroxylation of camphor to hydroxycamphor by the three- component P450cam system was chosen as the model reaction, and regeneration of NADH was accomplished with yeast alcohol dehydrogenase (YADH). P450cam was activated in the surfactant-free emulsions, and addition of AOT improved the activity even further, at least over the range of camphor concentrations for which initial rates were readily measurable in all media. The largest observed rate enhancement was 4.5-fold. Nearly 50-times more product was formed in the surfactant-stabilized emulsions than was achieved in aqueous buffer, with total turnover numbers reaching 28,900 for P450cam and 11,800 for YADH. In the absence of surfactant, the two-phase reaction appeared to be mass-transfer limited, while inclusion of AOT alleviated transport limitations and/or afforded a larger interfacial area for P450 activation. The oxidation of hydroxycamphor to 2,5-diketocamphane was also observed, owing to the large concentration of hydroxycamphor relative to camphor in the aqueous phase of the two-phase emulsion. This competing reaction was accompanied by the uncoupled oxidation of NADH (i.e., NADH oxidation without formation of 2,5-diketocamphane), which reduced the availability of NADH for camphor oxidation and further limited the yield of hydroxycamphor in the two-phase emulsions. These results indicate that a surfactant-stabilized two-phase emulsion is a promising reaction medium for practical P450 biocatalysis, although its effectiveness for a given P450/substrate combination can depend on several factors, including competitive or sequential reactions, product inhibition, and NAD(P)H uncoupling.  相似文献   

7.
Previous studies on mammalian peroxidases and cytochrome P450 family 4 enzymes have shown that a carboxylic group positioned close to a methyl group of the prosthetic heme is required for the formation of a covalent link between a protein carboxylic acid side chain and the heme. To determine whether there are additional requirements for covalent bond formation in the P450 enzymes, a glutamic acid or an aspartic acid has been introduced into P450(cam) close to the heme 5-methyl group. Spectroscopic and kinetic studies of the resulting G248E and G248D mutants suggest that the carboxylate group coordinates with the heme iron atom, as reported for a comparable P450(BM3) mutant [Girvan, H. M., Marshall, K. R., Lawson, R. J., Leys, D., Joyce, M. G., Clarkson, J., Smith, W. E., Cheesman, M. R., and Munro, A. W. (2004) J. Biol. Chem. 279, 23274-23286]. The two P450(cam) mutants have low catalytic activity, but in contrast to the P450(BM3) mutant, incubation of the G248E (but not G248D) mutant with camphor, putidaredoxin, putidaredoxin reductase, and NADH results in partial covalent binding of the heme to the protein. No covalent attachment is observed in the absence of camphor or any of the other reaction components. Pronase digestion of the G248E P450(cam) mutant after covalent attachment of the heme releases 5-hydroxyheme, establishing that the heme is covalently attached through its 5-methyl group as predicted by in silico modeling. The results establish that a properly positioned carboxyl group is the sole requirement for autocatalytic formation of a heme-protein link in P450 enzymes, but also show that efficient covalent binding requires placement of the carboxyl close to the methyl but in a manner that prevents strong coordination to the iron atom.  相似文献   

8.
Addition of alcohols to cytochrome P450cam (CYP101) was shown to release the substrate camphor from the heme pocket of the enzyme. The release of the substrate was found to be caused both due to increased solubility of the substrate in solution in presence of alcohol and due to change in the tertiary structure of the active site of the enzyme. The far-UV CD and near-UV CD spectra reveal that addition of alcohols to cytochrome P450cam cause a small change in the secondary structural elements but a significant change in the tertiary structural organization of this enzyme. The CD spectra at the heme region at various concentrations of alcohols indicate a substantial change in the tertiary structural organization around the heme moiety too. The equilibrium constant associated with the binding of camphor to Cyt P450cam is strongly dependent on the concentration of alcohols and the corresponding free energy associated with the binding is found to scale linearly with the concentration of alcohols. Kinetic experiments on binding of camphor to Cyt P450cam show that both k(on) and k(off) rate constants are strongly affected by addition of alcohols suggesting that alcohol expel camphor out of the heme cavity of Cyt P450cam by affecting tertiary structure of Cyt P450cam as well as by modifying the solubility properties of camphor in aqueous medium.  相似文献   

9.
It is generally assumed that the putative compound I (cpd I) in cytochrome P450 should contain the same electron and spin distribution as is observed for cpd I of peroxidases and catalases and many synthetic cpd I analogues. In these systems one oxidation equivalent resides on the Fe(IV)=O unit (d(4), S=1) and one is located on the porphyrin (S'=1/2), constituting a magnetically coupled ferryl iron-oxo porphyrin pi-cation radical system. However, this laboratory has recently reported detection of a ferryl iron (S=1) and a tyrosyl radical (S'=1/2), via M?ssbauer and EPR studies of 8 ms-reaction intermediates of substrate-free P450cam from Pseudomonas putida, prepared by a freeze-quench method using peroxyacetic acid as the oxidizing agent [Schünemann et al., FEBS Lett. 479 (2000) 149]. In the present study we show that under the same reaction conditions, but in the presence of the substrate camphor, only trace amounts of the tyrosine radical are formed and no Fe(IV) is detectable. We conclude that camphor restricts the access of the heme pocket by peroxyacetic acid. This conclusion is supported by the additional finding that binding of camphor and metyrapone inhibit heme bleaching at room temperature and longer reaction times, forming only trace amounts of 5-hydroxy-camphor, the hydroxylation product of camphor, during peroxyacetic acid oxidation. As a control we performed freeze-quench experiments with chloroperoxidase from Caldariomyces fumago using peroxyacetic acid under the identical conditions used for the substrate-free P450cam oxidations. We were able to confirm earlier findings [Rutter et al., Biochemistry 23 (1984) 6809], that an antiferromagnetically coupled Fe(IV)=O porphyrin pi-cation radical system is formed. We conclude that CPO and P450 behave differently when reacting with peracids during an 8-ms reaction time. In P450cam the formation of Fe(IV) is accompanied by the formation of a tyrosine radical, whereas in CPO Fe(IV) formation is accompanied by the formation of a porphyrin radical.  相似文献   

10.
RS-8359, (+/-)-4-(4-cyanoanilino)-5,6-dihydro-7-hydroxy-7H-cyclopenta[d]-pyrimidine is a racemic compound with a selective and reversible monoamine oxidase A (MAO-A) inhibition activity. The substrate and product enantioselectivity with respect to 2-hydroxylation of RS-8359 enantiomers was studied using mouse and rat liver microsomes. In mice, the (S)-enantiomer was transformed to the cis-diol metabolite, whereas the (R)-enantiomer to the trans-diol metabolite. The Vmax/Km value for the formation of the cis-diol metabolite from the (S)-enantiomer was sevenfold greater than that for the formation of the trans-diol metabolite from the (R)-enantiomer. The greater Vmax/Km value for the (S)-enantiomer was due to the tenfold smaller Km value compared to that for the (R)-enantiomer. The results were in fair agreement with the previously reported low plasma concentrations of the (S)-enantiomer and the high recovery of the cis-diol metabolite derived from the (S)-enantiomer in urine after oral administration of RS-8359 to mice. Similarly to mice, in rats the (R)-enantiomer was transformed to the trans-diol metabolite, whereas the (S)-enantiomer yielded the cis-diol and trans-diol metabolites. The Vmax/Km value for the (R)-enantiomer was larger than that for the (S)-enantiomer in rats, indicating that the low plasma concentration of the (S)-enantiomer in rats might be caused by a metabolic reaction other than P450-dependent hydroxylation. CYP3A was shown to be responsible for the trans-diol formation from the (R)-enantiomer.  相似文献   

11.
RS-8359, (+/-)-4-(4-cyanoanilino)-5,6-dihydro-7-hydroxy-7H-cyclopenta[d]pyrimidine selectively and reversibly inhibits monoamine oxidase A (MAO-A). After oral administration of rac-RS-8359 to rats, mice, dogs, monkeys, and humans, plasma concentrations of the (R)-enantiomer were greatly higher than were those of the (S)-enantiomer in all species studied. The AUC((R)) to AUC((S)) ratios were 2.6 in rats, 3.8 in mice, 31 in dogs, and 238 in monkeys, and the (S)-enantiomer was almost negligible in human plasma. After intravenous administration of RS-8359 enantiomers to rats, the pharmacokinetic parameters showed that the (S)-enantiomer had a 2.7-fold greater total clearance (CL(t)) and a 70% shorter half-life (t(1/2)) than those for the (R)-enantiomer but had no difference in distribution volume (V(d)). No significant difference in the intestinal absorption rate was observed. The principal metabolites were the 2-keto form, possibly produced by aldehyde oxidase, the cis-diol form, and the 2-keto-cis-diol form produced by cytochrome P450 in rats, the cis-diol form in mice, RS-8359 glucuronide in dogs, and the 2-keto form in monkeys and humans. Thus, the rapid disappearance of the (S)-enantiomer from the plasma was thought to be due to the rapid metabolism of the (S)-enantiomer by different drug-metabolizing enzymes, depending on species.  相似文献   

12.
The stereoselectivity of purified rat GSH transferases towards alpha-bromoisovaleric acid (BI) and its amide derivative alpha-bromoisovalerylurea (BIU) was investigated. GSH transferase 2-2 was the only enzyme to catalyse the conjugation of BI and was selective for the (S)-enantiomer. The conjugation of (R)- and (S)-BIU was catalysed by the isoenzymes 2-2, 3-3 and 4-4. Transferase 1-1 was less active, and no catalytic activity was observed with transferase 7-7. Isoenzymes 1-1 and 2-2 of the Alpha multigene family preferentially catalysed the conjugation of the (S)-enantiomer of BIU (and BI), whereas isoenzymes 3-3 and 4-4 of the Mu multigene family preferred (R)-BIU. The opposite stereoselectivity of conjugation of BI and BIU previously observed in isolated rat hepatocytes and the summation of activities of enzymes known to be present in hepatocytes on the basis of present data are in accord.  相似文献   

13.
The significance of disturbances of lipid metabolism caused by xenobiotic acyl-CoAs as possible causes of peroxisomal proliferation has been studied with the enantiomers of 2-phenylpropionic acid (2-PPA), the (R)-enantiomer of which is converted to the acyl-CoA in rats while its (S)-antipode is not. rac-2-PPA (250 mg/kg/day ip × 3) was shown to be an hepatic peroxisomal proliferator in male Sprague–Dawley rats on the basis of increases in microsomal cytochrome P-450 content and lauric acid hydroxylation and hepatic CN?-insensitive palmitoyl-CoA oxidation, a peroxisomal marker activity, while electron microscopy revealed a rise in the peroxisome/mitochondria ratio in hepatocytes. Further studies established the dose–response relationships for these biochemical changes. The (R)- and (S)-enantiomers were administered at a dose of 50 mg/kg/day ip × 3 and both were peroxisome proliferators of very similar potency. The effects of 100 mg/kg/day ip × 3 of the racemate, a dose giving ca. 75% of maximal response, were essentially additive of those of 50 mg/kg/day ip × 3 of its two component isomers. The stereoselectivity of acyl-CoA formation from the enantiomers of 2-PPA was confirmed by their differential inhibition of microsomal palmitoyl-CoA synthesis. Taken together, these data indicate that it is very unlikely that the acyl-CoA of 2-PPA plays any role in the peroxisomal proliferation which this compound causes in the rat. © 1994 Wiley-Liss, Inc.  相似文献   

14.
The principal oxidative metabolites formed from benzo(c)phenanthrene (B(c)Ph) by the cytochromes P450 in liver microsomes from control and treated rats are the 3,4- and 5,6-arene oxides. A procedure is described which allows determination of the enantiomer composition and absolute configuration of these arene oxides based on HPLC separation of isomeric thiolate adducts formed with N-acetyl-L-cysteine in base. Incubation of [3H]-B(c)Ph with highly purified cytochrome P450c in a reconstituted monooxygenase system followed by trapping of the metabolically formed arene oxides as above indicated that the 3,4-oxide was predominantly the (+)-(3S,4R)-enantiomer (90%) and that the 5,6-oxide consisted mainly of the (+)-(5S,6R)-enantiomer (76%). The results are discussed in terms of their implications about the catalytic binding site of cytochrome P450c.  相似文献   

15.
Cytochrome P450cam (P450CIA1) catalyzes the hydroxylation of camphor and several substrate analogues such as norcamphor and 1-methyl-norcamphor. Hydroxylation was found experimentally at the 3, 5, and 6 positions of norcamphor, but only at the 5 and 6 positions of 1-methyl-norcamphor. In the catalytic cycle, the hydroxylation of substrate is coupled to the consumption of NADH. For camphor, the degree of coupling is 100%, but for both norcamphor and 1-methyl-norcamphor, the efficiency is dramatically lowered to 12% and 50%, respectively. Based on an examination of the active site of P450cam, it appeared that mutating position 185 might dramatically alter the product specificity and coupling of hydroxylation of norcamphor by P450cam. Analysis of molecular dynamics trajectories of norcamphor bound to the T185F mutant of cytochrome P450cam predicted that hydroxylation at the 3 position should be abolished and that the coupling should be dramatically increased. This mutant was constructed and the product profile and coupling experimentally determined. The coupling was doubled, and hydroxylation at the 3 position was essentially abolished. Both of these results are in agreement with the prediction.  相似文献   

16.
P450cam has long served as a prototype for the cytochrome P450 (CYP) gene family. But, little is known about how substrate enters its active site pocket, and how access is achieved in a way that minimizes exposure of the reactive heme. We hypothesize that P450cam may first bind substrate transiently near the mobile F-G helix that covers the active site pocket. Such a two-step binding process is kinetically required if P450cam rarely populates an open conformation-as suggested by previous literature and the inability to obtain a crystal structure of P450cam in an open conformation. Such a mechanism would minimize exposure of the heme by allowing P450cam to stay in a closed conformation as long as possible, since only brief flexing into an open conformation would be required to allow substrate entry. To test this model, we have attempted to dock a second camphor molecule into the crystal structure of camphor-bound P450cam. The docking identified only one potential entry site pocket, a well-defined cavity on the F-helix side of the F-G flap, 16 A from the heme iron. Location of this entry site pocket is consistent with our NMR T1 relaxation-based measurements of distances for a camphor that binds in fast exchange (active site camphor is known to bind in slow exchange). Presence of a second camphor binding site is also confirmed with [(1)H-(13)C] HSQC titrations of (13)CH3-threonine labeled P450cam. To confirm that camphor can bind outside of the active site pocket, (13)CH3-S-pyridine was bound to the heme iron to physically block the active site, and to serve as an NMR chemical shift probe. Titration of this P450cam-pyridine complex confirms that camphor can bind to a site outside the active site pocket, with an estimated Kd of 43 microM. The two-site binding model that is proposed based on these data is analogous to that recently proposed for CYP3A4, and is consistent with recent crystal structures of P450cam bound to tethered-substrates, which force a partially opened conformation.  相似文献   

17.
Prasad S  Mitra S 《Biochemistry》2002,41(49):14499-14508
The role of protein structural flexibility and substrate dynamics in catalysis by cytochrome P450 enzymes is an area of current interest. We have addressed these in cytochrome P450(cam) (P450(cam)) and its Y96A mutant with camphor and its related compounds using fluorescence spectroscopy. Previously [Prasad et al. (2000) FEBS Lett. 477, 157-160], we provided experimental support to dynamic fluctuations in P450(cam), and substrate access into the active site region via the channel next to the flexible F-G helix-loop-helix segment. In the investigation described here, we show that the dynamic fluctuations in the enzyme are substrate dependent as reflected by tryptophan fluorescence quenching experiments. The orientation of tryptophan relative to heme (kappa(2)) for W42 obtained from time-resolved tryptophan fluorescence measurements show variation with type of substrate bound to P450(cam) suggesting regions distant from heme-binding site are affected by physicochemical and steric characteristics/protein-substrate interactions of P450(cam) active site. We monitored substrate dynamics in the active site region of P450(cam) by time-resolved substrate anisotropy measurements. The anisotropy decay of substrates bound to P450(cam) indicate that mobility of substrates is modulated by physicochemical and steric characteristics/protein-substrate interactions of local active site structure, and provides an understanding of factors controlling observed hydroxylated products for substrate bound P450(cam) complexes. The present study shows that P450(cam) local and peripheral structural flexibility and heterogeneity along with substrate mobility play an important role in regulating substrate binding orientation during catalysis and accommodating diverse range of substrates within P450(cam) heme pocket.  相似文献   

18.
P450BM-3, a catalytically self-sufficient, soluble bacterial P450, contains on the same polypeptide a heme domain and a reductase domain. P450BM-3 catalyzes the oxidation of short- and long-chain, saturated and unsaturated fatty acids. The three-dimensional structure of the heme domain both in the absence and in the presence of fatty acid substrates has been determined; however, the fatty acid in the substrate-bound form is not adequately close to the heme iron to permit a prediction regarding the stereoselectivity of oxidation. In the case of long-chain fatty acids, the products can also serve as substrate and be metabolized several times. In the current study, we have determined the absolute configuration of the three primary products of palmitic acid hydroxylation (15-, 14-, and 13-OH palmitic acid). While the 15- and 14-hydroxy compounds are produced in a highly stereoselective manner (98% R, 2% S), the 13-hydroxy is a mixture of 72% R and 28% S. We have also examined the binding of these three hydroxy acids to P450BM-3 and shown that only two of them (14-OH and 13-OH palmitic acid) can bind to and be further metabolized by P450BM-3. The results indicate that in contrast to the flexibility of palmitoleic acid bound to the oxidized enzyme, palmitic acid is rigidly bound in the active site during catalytic turnover.  相似文献   

19.
The stereoselectivity of the teratogenic activity of 2-ethylhexanoic acid (EHXA), a metabolite of the widely-used plasticizer di-(2-ethylhexyl)phthalate, was investigated. The enantiomers of EHXA were prepared via asymmetric synthesis with the aid of the chiral auxiliaries (R)- and (S)-1-amino-2-(methoxymethyl)pyrrolidine (RAMP, SAMP). The aqueous solutions of the sodium salts of (R)- and (S)-EHXA and the racemic EHXA [+/- )-EHXA) were injected each morning and evening of day 7 and 8 of gestation in the NMRI mouse (500 mg/kg, i.p.), a period highly sensitive in regard to the production of neural tube defects (exencephaly) by branched-chain carboxylic acids. (S)-EHXA did not yield any teratogenic or embryotoxic response in this model, while (R)-EHXA was highly teratogenic (59% of living fetuses exhibited exencephaly) and embryotoxic (as indicated by embryolethality and fetal weight retardation); the exencephaly rate induced by (+/- )-EHXA was between those of the two enantiomeres (32%). It is therefore likely that stereoselective interactions of the enantiomers of EHXA with chiral molecules in the embryo are decisive in regard to the teratogenic response. This first example of the stereoselectivity of the teratological activity of an environmental pollutant suggests that the safety of man-made chemicals can be improved by the use of pure enantiomers instead of racemates.  相似文献   

20.
Male C57BL/6 mice were exposed to 1% (w/w) (+)- or (?)-2-ethylhexanoic acid or an equimolar mixture of these enantiomers in their diet for 4 or 10 days. A significant increase in liver weight and a 2- to 3-fold increase in the protein content of the mitochondrial fraction were seen in all cases. Peroxisomal palmitoyl-CoA oxidation was increased 2- to 3.5-fold after 4 days of treatment and 4- to 5-fold after 10 days, while the corresponding increases in peroxisomal lauroyl-CoA oxidase activity were 2- to 3-fold and 9- to 12-fold, respectively. Peroxisomal catalase activity was unchanged, whereas the microsomal and cytosolic activities were increased 2- to 3-fold and 6- to 16-fold, respectively. These treatments also induced microsomal ω-hydroxylation of lauric acid 7-fold and soluble epoxide hydrolase activity in the mitochondrial and cytosolic fractions, as well as microsomal epoxide hydrolase activity about 50–100%. The only significant differences observed between the effects of (+)-2-ethylhexanoic acid and its (?)-enantiomer were on peroxisomal palmitoyl-CoA oxidation and lauroyl-CoA oxidase activity after 4 days of treatment. In both these cases the (+)-enantiomer resulted in increases which were 50–75% greater than those seen with the (?)-form. © 1994 Wiley-Liss, Inc.  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