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.
Capsaicin and its analogs with different acyl moieties were found to inhibit the electron-transfer activity of NADH-coenzyme Q oxidoreductase isolated from beef heart mitochondria. The inhibitory potency of capsaicin was lower than those of dihydrocapsaicin and analogs with heptanoyl, capryl, undecanoyl, and lauroyl moieties, but was higher than those of analogs with palmitoyl and stearoyl moieties. The analog with the lauroyl moiety showed the strongest inhibition. These results suggest that hydrophobicity and the appropriate carbon chain length of the acyl moiety are important for the binding of compounds to the enzyme. On the other hand, capsaicin and its analogs did not interrupt rotenone-insensitive electron transfer from NADH to menadione. Furthermore, these compounds had almost no effect on the spectral properties and EPR signals arising from iron-sulfur clusters of the NADH-treated enzyme. Kinetic analyses with double-reciprocal plots showed that these compounds were competitive inhibitors with respect to coenzyme Q1, an electron acceptor. These results strongly suggest that capsaicin and its analogs bind to the coenzyme Q1 binding site of the enzyme.  相似文献   

2.
Evidence for coenzyme Q function in transplasma membrane electron transport   总被引:2,自引:0,他引:2  
Transplasma membrane electron transport activity has been associated with stimulation of cell growth. Coenzyme Q is present in plasma membranes and because of its lipid solubility would be a logical carrier to transport electrons across the plasma membrane. Extraction of coenzyme Q from isolated rat liver plasma membranes decreases the NADH ferricyanide reductase and added coenzyme Q10 restores the activity. Piericidin and other analogs of coenzyme Q inhibit transplasma membrane electron transport as measured by ferricyanide reduction by intact cells and NADH ferricyanide reduction by isolated plasma membranes. The inhibition by the analogs is reversed by added coenzyme Q10. Thus, coenzyme Q in plasma membrane may act as a transmembrane electron carrier for the redox system which has been shown to control cell growth.  相似文献   

3.
The biological activities of tamoxifen derivatives that contain various side chain alterations were studied using a T47D breast cancer cell growth assay in vitro. We studied the activity of various analogs to determine the important aspects of side chain composition and aryl ring positioning on antiestrogenic activity. Previous studies utilizing a rat pituitary cell prolactin synthesis assay have shown that substitution of the aminoethoxy side chain for an allyl side chain resulted in agonist activity, whereas the addition of a glyceryl side chain produced antiestrogenic activity. In the present study utilizing T47D cells, compounds with alkyl or allyl substitutions were partial agonists, as were compounds with bulky para-substituted benzyl group constituents. A tamoxifen derivative with a side chain containing an ethyl ester was antiestrogenic (IC50 = 2 x 10(-6) M) and effectively inhibited estradiol (10(-10) M) stimulation of growth. However, a compound with a short similar methyl ester-containing side chain did not possess any activity. Compounds with carbinol-containing side chains were antiestrogenic (IC50 = 2.8-3.5 x 10(-7) M). All of the compounds displaying antiestrogenic activity could be "rescued" by incubation with estradiol (10(-8) M) and therefore were not nonspecifically toxic to the cells. These results support the hypothesis that the presence of a lone pair of electrons within the side chain region of tamoxifen may be required for antiestrogenic activity. Also, nonplanar placement of the aryl ring of the triphenylethylene-type of compound is critical for potency.  相似文献   

4.
In earlier work, we synthesized a cyclic 9-amino acid peptide (AFPep, cyclo[EKTOVNOGN]) and showed it to be useful for prevention and therapy of breast cancer. In an effort to explore the structure–function relationships of AFPep, we have designed analogs that bear a short ‘tail’ (one or two amino acids) attached to the cyclic peptide distal to its pharmacophore. Analogs that bore a tail of either one or two amino acids, either of which had a hydrophilic moiety in the side chain (e.g., cyclo[EKTOVNOGN]FS) exhibited greatly diminished biological activity (inhibition of estrogen-stimulated uterine growth) relative to AFPep. Analogs that bore a tail of either one or two amino acids which had hydrophobic (aliphatic or aromatic) side chains (e.g., cyclo[EKTOVNOGN]FI) retained (or had enhanced) growth inhibition activity. Combining in the same biological assay a hydrophilic-tailed analog with either AFPep or a hydrophobic-tailed analog resulted in decreased activity relative to that for AFPep or for the hydrophobic-tailed analog alone, suggesting that hydrophilic-tailed analogs are binding to a biologically active receptor. An analog with a disrupted pharmacophore (cyclo[EKTOVGOGN]) exhibited little or no growth inhibition activity. An analog with a hydrophilic tail and a disrupted pharmacophore (cyclo[EKTOVGOGN]FS) exhibited no growth inhibition activity of its own and did not affect the activity of a hydrophobic-tailed analog, but enhanced the growth inhibition activity of AFPep. These results are discussed in the context of a two-receptor model for binding of AFPep and ring-and-tail analogs. We suggest that tails on cyclic peptides may comprise a useful method to enhance diversity of peptide design and specificity of ligand–receptor interactions.  相似文献   

5.
To explore the hydrophobic groove subsite within the CB1 cannabinoid receptor we have designed and synthesized a group of tail-substituted anandamide analogs. Our design involves the introduction of aryl or heterocyclic ring as terminal substituents that are connected to the last cis-arachidonyl double bond through aliphatic chains of variable lengths. Our results indicate that there are strict stereochemical requirements for the interaction of such analogs with the CB1 receptor. The optimal pharmacophore includes the phenyl, p-substituted phenyl, or 3-furyl substituents attached to the cis-double bond through a four methylene chain.  相似文献   

6.
In this study, we used molecules with either of the structural differences in the side chains of vitamin D(2) and vitamin D(3) to investigate which feature is responsible for the significant differences in their respective metabolism, pharmacokinetics and toxicity. We used two cell model systems-HepG2 and HPK1A-ras-to study hepatic and target cell metabolism, respectively. Studies with HepG2 revealed that the pattern of 24- and 26-hydroxylation of the side chain reported for 1alpha-hydroxyvitamin D(2) (1alpha-OH-D(2)) but not for 1alpha-OH-D(3) is also observed in both 1alpha-OH-D(4) and Delta(22)-1alpha-OH-D(3) metabolism. This suggests that the structural feature responsible for targeting the enzyme to the C24 or C26 site could be either the C24 methyl group or the 22-23 double bond. In HPK1A-ras cells, the pattern of metabolism observed for the 24-methylated derivative, 1alpha,25-(OH)(2)D(4), was the same pattern of multiple hydroxylations at C24, C26 and C28 seen for vitamin D(2) compounds without evidence of side chain cleavage observed for vitamin D(3) derivatives, suggesting that the C24 methyl group plays a major role in this difference in target cell metabolism of D(2) and D(3) compounds. Novel vitamin D(4) compounds were tested and found to be active in a variety of in vitro biological assays. We conclude that vitamin D(4) analogs and their metabolites offer valuable insights into vitamin D analog design, metabolic enzymes and maybe useful clinically.  相似文献   

7.
The discovery and the rediscovery of plastoquinone (PQ) are described together with the definition of its structure as a 2,3-dimethyl 5 solanosyl benzoquinone. The discovery, by M. Kofler, was a result of a search for Vitamin K. Its rediscovery was made by me, when I was at The Enzyme Institute of the University of Wisconsin, analyzing animals and plants for the newly discovered coenzyme Q. In green plants, I found another lipophilic quinone in addition to coenzyme Q. Some misleading evidence suggested as if the new quinone had coenzyme Q activity in mitochondria, but improved methods gave negative results. When I found that the quinone was concentrated in chloroplasts, I considered a role for it in photosynthesis analogous to the role of coenzyme Q in mitochondria. After moving to the Chemistry Department, University of Texas at Austin, I used a plain light bulb and some spinach chloroplasts to show that PQ could be involved in photosynthetic redox reactions. This effect was supported by Norman Bishop’s restoration of chloroplast electron transport after solvent extraction, with PQ and photoreduction studies by E. R. Redfern and J. Friend in R. A. Morton’s laboratory in Liverpool, UK. We also found an additional analog of PQ in addition to a second analog found in Wisconsin. We called the new analogs PQB and PQC. Although we found some restoration effects with PQC, the discovery by W. T. Griffiths in Morton’s laboratory, that PQB and PQC consisted of six forms of PQ each, made it more likely that the new analogs were breakdown products. Morton’s group established the structure of the PQCs as a series of PQs, with a hydroxyl group on the prenyl side chain, and the PQB series as having fatty acids esterified to the hydroxyl groups of PQC. Possible functions of the analogs are also discussed in this article.  相似文献   

8.
Nigrostriatal cell death in 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-induced Parkinson's disease results from the inhibition of mitochondrial respiration by 1-methyl-4-phenylpyridinium (MPP+). MPP+ blocks electron flow from NADH dehydrogenase to coenzyme Q at or near the same site as do rotenone and piericidin and protects against binding of and loss of activity due to these inhibitors. The 4'-analogs of MPP+ showed increasing affinity for the site with increasing length of alkyl chain, with the lowest Ki, for 4'-heptyl-MPP+, being 6 microM. The 4'-analogs compete with rotenone for the binding site in a concentration-dependent manner. They protect the activity of the enzyme from inhibition by piericidin in parallel to preventing its binding, indicating that the analogs and piericidin bind at the same inhibitory site(s). The optimum protection, however, was afforded by 4'-propyl-MPP+. The lesser protection by the more lipophilic MPP+ analogs with longer alkyl chains may involve a different orientation in the hydrophobic cleft, allowing rotenone and piericidin to still bind even when the pyridinium cation is in a position to interrupt electron flow from NADH to coenzyme Q.  相似文献   

9.
Gramicidin S (GS) analogs in which the Ndelta atoms of the two Orn side chains are linked by an oligomethylene bridge [-(CH2)n-; n=3-5] were prepared via the bis(p-nitrobenzenesulfonyl) derivative [Orn(NBS)2,2']GS. For comparison the nonbridged secondary amino group-containing analog [Orn(Me)2,2']GS was also prepared. 1H NMR and CD spectral analysis indicated that these analogs adopt the same beta-sheet conformation as GS. The antimicrobial activities of these analogs were very similar, but were slightly dependent on the bridge chain length, the trimethylene-bridged analog being the most potent.  相似文献   

10.
Germination of Clostridium difficile spores is the first required step in establishing C. difficile-associated disease (CDAD). Taurocholate (a bile salt) and glycine (an amino acid) have been shown to be important germinants of C. difficile spores. In the present study, we tested a series of glycine and taurocholate analogs for the ability to induce or inhibit C. difficile spore germination. Testing of glycine analogs revealed that both the carboxy and amino groups are important epitopes for recognition and that the glycine binding site can accommodate compounds with more widely separated termini. The C. difficile germination machinery also recognizes other hydrophobic amino acids. In general, linear alkyl side chains are better activators of spore germination than their branched analogs. However, L-phenylalanine and L-arginine are also good germinants and are probably recognized by distinct binding sites. Testing of taurocholate analogs revealed that the 12-hydroxyl group of taurocholate is necessary, but not sufficient, to activate spore germination. In contrast, the 6- and 7-hydroxyl groups are required for inhibition of C. difficile spore germination. Similarly, C. difficile spores are able to detect taurocholate analogs with shorter, but not longer, alkyl amino sulfonic acid side chains. Furthermore, the sulfonic acid group can be partially substituted with other acidic groups. Finally, a taurocholate analog with an m-aminobenzenesulfonic acid side chain is a strong inhibitor of C. difficile spore germination. In conclusion, C. difficile spores recognize both amino acids and taurocholate through multiple interactions that are required to bind the germinants and/or activate the germination machinery.  相似文献   

11.
ABA analog structure-function relationships were determined by testing an array of 19 different ABA analogs on 1-year-old clonal white spruce ( Picea glauca [Moench.] Voss) raised from somatic embryos. The contribution of specific structural features to analog activity was determined from the relative effect of aeroponically applied analog solutions (10−3 M ) on seedling gas exchange. Seedling transpiration rate (E) and carbon assimilation rate (A) were measured continuously during treatment by means of a whole plant cuvette system. The analogs were racemic about the C-1' chiral center and were derived from changes imposed on six regions of the ABA molecule. The activity of optically pure (+)-S-ABA and (−)-R-ABA were also determined. Analog activity was reduced by changing the oxidation level at C-1 from the carboxylic acid. The ring C-2', C-3' double bond was important but not essential to activity. The activity lost through changes in ring structure and C-1 oxidation level was, in many cases, almost fully restored by replacing the C-4, C-5 double bond with a triple bond. Therefore, analogs with a triple bond at C-4 were more active than their equivalents with a dienoic side chain. Fluorination of the C-7' methyl caused a relatively moderate reduction in analog activity. Truncation of C-1 and C-2 from the side chain reduced activity to near zero. The unnatural (−)-ABA enantiomer was inactive.  相似文献   

12.
In a protein synthesis sytem derived from porcine uteri we have made the following observations: 1. Synthesis directed by the endogenous mRNA of the system is more sensitive to inhibition by puromycin than poly(U) directed synthesis. 2. Synthesis directed by the synthetic templates poly(G,U) and poly(C,U) is more sensitive to inhibition by puromycin than poly(U) directed synthesis and frequently shows a sensitivity to puromycin similar to that directed by endogenous mRNA. 3. Protein synthesis was inhibited by three aminoacyloligonucleotides (C-A-Phe, C-A-Asp, and C-A-Pro) which are analogs of the 3' terminus of aminoacyl tRNAs. Of the three, C-A-Phe was the most active and C-A-Asp the least active but, unlike puromycin, inhibition by these compounds was the same for endogenous and poly(U) directed synthesis. These results are interpreted as supporting the proposal that the acceptor site of ribosomes contains an hydrophobic binding region which interacts with the side chain of aliphatic amino acids, and particularly with the aromatic side chain of phenylalanine.  相似文献   

13.
Synthetic analogs of (2'-5')oligo(A) were assayed for endonuclease activation in cell extracts and for inhibition of protein synthesis in intact cells. The analogs are triadenylates: (i) methylated in the terminal 3'-OH; (ii) methylated at all three 3'-OH groups; (iii) with different numbers of phosphate groups at the 5' terminus or with a methylene group between the beta- and gamma-phosphate. Only 5'-phosphorylated monomethylated analogs activate an endonuclease in cell extracts and are powerful inhibitors of protein synthesis in intact cells. The analogs with only one 5'-terminal phosphate may require addition of another phosphate for activity since the kinase inhibitor 2-aminopurine prevents endonuclease activation by this compound but not by the di- and triphosphate-terminated triadenylates. These results suggest that two terminal phosphates and one or two free 3'-OH are required for endonuclease activation and inhibition of protein synthesis. The monomethylated analogs are more active than (2'-5')pppA3 because of their resistance to degradation by cellular enzymes. Accordingly, the monomethylated analogs cause a prolonged inhibition of protein synthesis in human fibroblasts treated with nanomolar concentrations of these compounds.  相似文献   

14.
We recently found that a thioether analog of K vitamin (Cpd 5) inhibited the activity of protein-tyrosine phosphatases (PTPases) and induced protein-tyrosine phosphorylation in a human hepatoma cell line (Hep3B). We have now examined the structural requirements for induction of protein-tyrosine phosphorylation and PTPase inhibition by several K vitamin analogs. Thioether analogs with sulfhydryl arylation capacity, especially those with a hydroxy (Cpd 5) or a methoxy group at the end of the side chain, induced protein-tyrosine phosphorylation, but non-arylating analogs, such as those with an all-carbon or O-ether side chain, did not. Among the receptor-tyrosine kinases, epidermal growth factor receptors were tyrosine-phosphorylated by treatment with thioether analogs, whereas insulin and hepatocyte growth factor receptors were not. An increase in tyrosine-phosphorylated ERK2 mitogen-activated protein kinase was also observed. The activity of purified T cell PTPase was inhibited only by the thioether analogs, but not by non-arylating analogs. Furthermore, the epidermal growth factor receptor dephosphorylation activity of Hep3B cell lysates was inhibited by Cpd 5 treatment. A similar induction of protein-tyrosine phosphorylation by Cpd 5 was seen in other human hepatoma cell lines together with growth inhibition. However, one cell line (HepG2), which was relatively resistant to growth inhibition by Cpd 5, did not increase its phosphorylation levels upon Cpd 5 treatment. These results suggest that cell growth inhibition by thioether analogs is closely associated with inhibition of PTPases by sulfhydryl arylation and with tyrosine phosphorylation of selected proteins.  相似文献   

15.
I. L. Sun  E. E. Sun  F. L. Crane 《Protoplasma》1995,184(1-4):214-219
Summary The addition of coenzyme Q10 to culture media stimulates the serum-free growth of HeLa, HL-60 cells, and mouse fibroblasts (Balb/3T3). With HeLa cells, the stimulation by coenzyme Q10 is additive to the stimulation by ferricyanide, an impermeable electron acceptor for the transplasma membrane electron transport. This combined response to coenzyme Q10 and ferricyanide is enhanced with insulin. -Tocopherylquinone can also stimulate the growth of HeLa cells, but vitamin K1 is inactive. Specificity of quinone effects is indicated. Serum-free growth of Balb/3T3 and SV 40 transformed BaIb/3T3 (SV/T2) cells is also stimulated by coenzyme Qio with stimulation similar to HeLa cells. However, Balb/3T3 cells are not stimulated by ferricyanide, which does not increase the response to coenzyme Q10. The transformed cells (SV/T2) respond better to ferricyanide alone, but the effects of coenzyme Qio and ferricyanide are not additive. Serum-free growth of HL-60 cells is stimulated dramatically by coenzyme Q10. The extent of growth stimulation on HL-60 cells is almost six-fold that of HeLa or Balb/3T3 cells. The stimulation of NADH-ferricyanide reductase (a transmembrane redox enzyme) by coenzyme Q10 with HL-60 cells is similar to their growth pattern in response to coenzyme Q10. Unlike HL-60, HeLa and Balb/3T3 cells show little stimulation of ferricyanide reduction by coenzyme Q10. The stimulatory effect on both ferricyanide reduction and cell growth by the short side-chain coenzyme Q2 is much less than that of the long side-chain coenzyme Q10. Ferricyanide reduction by HeLa cells is inhibited by coenzyme Q analogs such as 2,3-dimethoxy-5-chloro-6-naphthyl-mercapto-coenzyme Q and 2-methoxy-3-ethoxyl-5-methyl-6-hexadecyl-mercapto-coenzyme Q. However, these inhibitions are reversed by coenzyme Q10. The growth inhibition of HL-60 cells by other coenzyme Q analogs, such as capsiacin can also be reversed by coenzyme Q10. These data indicate that plasma membrane-based NADH oxidation or modification of the membrane quinone redox balance may be a basis for the growth stimulation.  相似文献   

16.
1. Dequalinium chloride (DECA) and three related quinolinium compounds inhibit bovine heart mitochondrial and Paracoccus denitrificans electron transport activity, with inhibition localized between NADH and ubiquinone in both electron transport chains. 2. Structure-activity studies reveal that two quinolinium rings and a long bridging group are necessary for significant inhibition of reduction of artificial electron acceptors and coenzyme Q, whereas only one quinolinium ring and a long hydrocarbon side chain are required for significant inhibition of NADH oxidase activity. 3. Inhibition of coenzyme Q reduction by DECA is not reversed by dialysis. 4. Studies comparing DECA inhibition of rotenone-sensitive with rotenone-insensitive preparations indicate that DECA acts by a different inhibitory mechanism than rotenone on mammalian mitochondrial and P. denitrificans NADH----ubiquinone reductase.  相似文献   

17.
Sphingolipids have been found to have single methyl branchings both in their long-chain base and in their N-linked acyl chains. In this study we determined how methyl-branching in the N-linked acyl chain of sphingomyelin (SM) affected their membrane properties. SM analogs with a single methyl-branching at carbon 15 (of a 17:0 acyl chain; anteiso) had a lower gel-liquid transition temperature as compared to an iso-branched SM analog. Phytanoyl SM (methyls at carbons 3, 7, 11 and 15) as well as a SM analog with a methyl on carbon 10 in a hexadecanoyl chain failed to show a gel-liquid transition above 10 °C. Only the two distally branched SM analogs (iso and anteiso) formed ordered domains with cholesterol in a 1-palmitoyl-2-oleoyl-phosphatidylcholine (POPC) bilayer. However, domains formed by the branched SM analogs appeared to contain less sterol when compared to palmitoyl SM (PSM) as the saturated phospholipid. Sterol-enriched domains formed by the anteiso SM analog were also less stable against temperature than domains formed by PSM. Both the 10-methyl and phytanoyl SM analogs failed to form sterol-enriched domains in the POPC bilayer. Acyl chain branching weakened SM/sterol interactions markedly when compared to PSM, as also evidenced from the decreased affinity of cholestatrienol to bilayers containing branched SM analogs. Our results show that methyl-branching weakened intermolecular interactions in a position-dependent manner.  相似文献   

18.
From the intact cells of “Rhodotorula lactosa” R1 (IFO 1058), a new coenzyme Q, which has a different mobility on paper chromatograms from other five naturally occurring homologs of the coenzyme Q series, was isolated and purified as a crystalline state. The chemical analyses such as UV and IR absorption spectrophotometries, and NMR and mass spectrometries revealed that the material, mp 28.7~28.9°C, was identified as a Co Q10 derivative with the reduced C5 unit in the isoprenoid side chain terminal remote from the quinone nucleus, Co Q10 (H–10). The strain R 1 with such a unique coenzyme Q system is, concerning its taxonomic position, discussed in connection with other criteria.  相似文献   

19.
The catalyzed coupling reaction of activated alcohol and mercaptan was used for the short and efficient synthesis of 14 thioether compounds. Two types of side chains, the methyl butyl alkylamide related to the pure steroidal antiestrogen ICI 164384 and the dimethylamino ethyloxy phenyl related to the clinically used nonsteroidal antiestrogen tamoxifen, were introduced by a thioether link on two types of nuclei (triphenylethane or estradiol). The new thioether derivatives were tested to assess their relative binding affinity for the estrogen receptor and their estrogenic or antiestrogenic activity in the ZR-75-1 (ER+) cell line. The results indicate that of the three types of compounds studied, only the nonsteroidal derivatives with an alkylamide side chain possess antiestrogenic activity. In the steroidal series, displacement of the alkylamide side chain from the 7 to the 6 position produced compounds with chemical characteristics similar to ICI 164384 or EM-139 but without antiestrogenic activity. In the nonsteroidal series of compounds with an aryl side chain, compounds with estrogenic activity were obtained. One compound, a nonsteroidal derivative with a methyl butyl alkylamide side chain 20, possesses a relative binding affinity for the estrogen receptor identical to EM-139 (1.1 and 1.2%, respectively) and a relatively good antiestrogenic activity that is 10-fold lower than EM-139 (IC50 values of 250 and 25 nM, respectively). This nonsteroidal thioether with an alkylamide side chain is free of estrogenic activity.  相似文献   

20.
N-Acetylserotonin (compound 1) and N-acetyldopamine (compound 7) inhibit bovine adrenal medullary sepiapterin reductase in a manner competitive with the pterin substrate and have Ki values of 0.12 and 0.4 microM, respectively. Molecular modeling suggests that the phenyl rings of the two compounds bind in the pyrimidine pocket of the enzyme with the 3-hydroxyl of dopamine or the 5-hydroxyl of serotonin aligned at the pyrimidine 4-position. Further, the acetyl moieties of the two inhibitors appear to mimic the substrate side chain. Consistent with this analysis, N-acetyl-m-tyramine (compound 13) is also an excellent competitive inhibitor (Ki = 0.13 microM), whereas N-acetyltryptamine (compound 2), N-acetyl-p-tyramine (compound 14) and N-acetylphenylethylamine (compound 15) all bind poorly. Interestingly, restricted-rotation analogs of N-acetyldopamine and N-acetyl-m-tyramine are noncompetitive inhibitors of the enzyme. Modification of N-acetyldopamine to N-chloroacetyldopamine (compound 10) or of N-acetylserotonin to the N-chloroacetyl (5) or N-methoxyacetyl (compound 6) analogs results in greatly increased competitive affinity, with Ki = 0.014 microM for the dopamine analog and 0.006 and 0.008 microM, respectively, for the serotonin analogs. In MOLT-4 T-cell leukemia and MCF-7 breast adenocarcinoma in culture, 0.1 mM N-methoxyacetylserotonin depleted tetrahydrobiopterin by greater than or equal to 97 and greater than 50%, respectively, with no effect upon cell growth. In both cell lines, the GTP cyclohydrolase inhibitor, 2,4-diamino-6-hydroxypyrimidine at 1-5 mM also depleted tetrahydrobiopterin greater than or equal to 97%. In this case, however, modest growth inhibition did occur. Since the growth inhibition could not be reversed upon tetrahydrobiopterin repletion, inhibition was due to other effects of the inhibitor rather than to tetrahydrobiopterin depletion. The results show that there is no effect on cell growth when at least 97% of the tetrahydrobiopterin in these cell lines is depleted. Since the sepiapterin reductase inhibitor depleted tetrahydrobiopterin with fewer nonspecific effects than the cyclohydrolase inhibitor, it will be useful for determining metabolic effects of tetrahydrobiopterin depletion.  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