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 359 毫秒
1.
The influence of side‐chain structure on the mode of reaction of ortho‐quinone amines has been investigated with a view, ultimately, to developing potential methods of therapeutic intervention by manipulating the early stages of melanogenesis. Four N‐substituted dopamine derivatives have been prepared and quinone formation studied using pulse radiolysis and tyrosinase‐oximetry. Ortho‐quinones with an amide or urea side chain were relatively stable, although evidence for slow formation of isomeric para‐quinomethanes was observed. A thiourea derivative cyclized fairly rapidly (k = 1.7/s) to a product containing a seven‐membered ring, whereas a related amidine gave more rapidly (k ~ 2.5 × 102/s) a stable spirocyclic product. The results suggest that cyclization of amides, ureas and carbamates (NHCO‐X; X = R, NHR or OR) does not occur and is not, therefore, a viable approach to the formation of tyrosinase‐activated antimelanoma prodrugs. It is also concluded that for N‐acetyldopamine spontaneous ortho‐quinone to para‐quinomethane isomerization is slow.  相似文献   

2.
trans‐Resveratrol (3,5,4′‐trihydroxy‐trans‐stilbene, RES), a naturally occurring polyphenol, has recently attracted increased interest as a health‐beneficial agent. However, based on its p‐substituted phenol structure, RES is expected to be a substrate for tyrosinase and to produce a toxic o‐quinone metabolite. The results of this study demonstrate that the oxidation of RES by tyrosinase produces 4‐(3′,5′‐dihydroxy‐trans‐styrenyl)‐1,2‐benzoquinone (RES‐quinone), which decays rapidly to an oligomeric product (RES‐oligomer). RES‐quinone was identified after reduction to its corresponding catechol, known as piceatannol. RES‐quinone reacts with N‐acetylcysteine, a small thiol, to form a diadduct and a triadduct, which were identified by NMR and MS analyses. The production of a triadduct is not common for o‐quinones, suggesting a high reactivity of RES‐quinone. RES‐quinone also binds to bovine serum albumin through its cysteine residue. RES‐oligomer can oxidize GSH to GSSG, indicating its pro‐oxidant activity. These results suggest that RES could be cytotoxic to melanocytes due to the binding of RES‐quinone to thiol proteins.  相似文献   

3.
RS‐4‐(4‐Hydroxyphenyl)‐2‐butanol (rhododendrol, RD) was reported to induce leukoderma of the skin. To explore the mechanism underlying that effect, we previously showed that oxidation of RD with mushroom tyrosinase produces RD‐quinone, which is converted to secondary quinone products, and we suggested that those quinones are cytotoxic because they bind to cellular proteins and produce reactive oxygen species. We then confirmed that human tyrosinase can oxidize both enantiomers of RD. In this study, we examined the metabolism of RD in B16F1 melanoma cells in vitro. Using 4‐amino‐3‐hydroxy‐n‐butylbenzene as a specific indicator, we detected moderate levels of RD‐pheomelanin in B16F1 cells exposed to 0.3 to 0.5 mM RD for 72 h. We also confirmed the covalent binding of RD‐quinone to non‐protein thiols and proteins through cysteinyl residues. The covalent binding of RD‐quinone to proteins was 20‐ to 30‐fold greater than dopaquinone. These results suggest that the tyrosinase‐induced metabolism of RD causes melanocyte toxicity.  相似文献   

4.
A series of novel isoflavonoids were synthesized based on structural modifications of daidzein, an active ingredient of traditional Chinese medicine (TCM) and evaluated for their anti‐influenza activity, in vitro, against H1N1 Tamiflu‐resistant (H1N1 TR) virus in the MDCK cell line. Among them, 4‐oxo‐4H‐1‐benzopyran‐8‐carbaldehydes 11a – 11g were most promising, and they demonstrated better activities and selectivities comparable to those the reference ribarivin, a nucleoside antiviral agent. 3‐(4‐Bromophenyl)‐7‐hydroxy‐4‐oxo‐4H‐1‐benzopyran‐8‐carboxaldehyde ( 11c ) displayed the best inhibitory activity (EC50, 29.0 μM ) and selectivity index (SI>10.3). Analysis of the structure? activity relationships (SAR) indicated that both the non‐naturally‐occurring Br‐substituted B‐ring and appropriate CHO and OH groups on the A‐ring might be critical for the activity and selectivity against H1N1 TR influenza viruses.  相似文献   

5.
Two new phenylpropanoids were isolated from Lindelofia stylosa (Kar . and Kir .) and characterized as 4‐hydroxy‐N‐{4‐[(E)‐3‐(4‐hydroxy‐3‐methoxyphenyl)prop‐2‐enamido]butyl}benzamide ( 1 ) and 2‐[3‐hydroxy‐4‐(4‐hydroxyphenoxy)phenyl]‐1‐(methoxycarbonyl)ethyl (E)‐3‐(3,4‐dihydroxyphenyl)prop‐2‐enoate ( 2 ). Four known compounds, i.e. two phenylpropanoids, p‐coumaric acid (=(E)‐3‐(4‐hydroxyphenyl)prop‐2‐enoic acid; 3 ) and ferulic acid (=(E)‐3‐(4‐hydroxy‐3‐methoxyphenyl)prop‐2‐enoic acid; 4 ), and two naphthalene glycosides, 8‐Oβ‐D ‐glucopyranosyltorachrysone ( 5 ) and 8‐Oβ‐D ‐glucopyranosyl‐6‐demethoxytorachrysone ( 6 ), were also isolated for the first time from the plant. Compounds 1 – 6 were subjected to various antioxidant assays, including DPPH radical‐ and superoxide anion‐scavenging, and Fe2+‐chelation assays. Compound 2 was found to be most active in all assays with potency nearly similar to that of propyl gallate. Besides 2 , compounds 1 and 5 were also found to be active in DPPH radical‐scavenging standard assay.  相似文献   

6.
A new polychlorinated triphenyl diether named microsphaerol ( 1 ), has been isolated from the endophtic fungus Microsphaeropsis sp. An intensive phytochemical investigation of the endophytic fungus Seimatosporium sp., led to the isolation of a new naphthalene derivative named seimatorone ( 2 ) and eight known compounds, i.e., 1‐(2,6‐dihydroxyphenyl)‐3‐hydroxybutan‐1‐one ( 3 ), 1‐(2,6‐dihydroxyphenyl)butan‐1‐one ( 4 ), 1‐(2‐hydroxy‐6‐methoxyphenyl)butan‐1‐one ( 5 ), 5‐hydroxy‐2‐methyl‐4H‐chromen‐4‐one ( 6 ), 2,3‐dihydro‐5‐hydroxy‐2‐methyl‐4H‐chromen‐4‐one ( 7 ), 8‐methoxynaphthalen‐1‐ol ( 8 ), nodulisporins A and B ( 9 and 10 , resp.), and daldinol ( 11 ). The structures of 1 and 2 were elucidated by detailed spectroscopic analysis including 1H‐ and 13C‐NMR, COSY, HMQC, HMBC, and HR‐EI‐MS, while the structures of the known compounds were deduced from comparison of their spectral data with those in the literature. Preliminary studies revealed that microsphaerol ( 1 ) showed good antibacterial activities against B. Megaterium and E. coli, and good antilagal and antifungal activities against C. fusca, M. violaceum, respectively. On the other hand, seimatorone ( 2 ) exhibited moderate antibacterial, antialgal, and antifungal activities.  相似文献   

7.
Structure? activity relationships of nine thiophenes, 2,2′: 5′,2″‐terthiophene ( 1 ), 2‐chloro‐4‐[5‐(penta‐1,3‐diyn‐1‐yl)thiophen‐2‐yl]but‐3‐yn‐1‐yl acetate ( 2 ), 4‐(2,2′‐bithiophen‐5‐yl)but‐3‐yne‐1,2‐diyl diacetate ( 3 ), 4‐[5‐(penta‐1,3‐diyn‐1‐yl)thiophen‐2‐yl]but‐3‐yne‐1,2‐diyl diacetate ( 4 ), 4‐(2,2′‐bithiophen‐5‐yl)‐2‐hydroxybut‐3‐yn‐1‐yl acetate ( 5 ), 2‐hydroxy‐4‐[5‐(penta‐1,3‐diyn‐1‐yl)thiophen‐2‐yl]but‐3‐yn‐1‐yl acetate ( 6 ), 1‐hydroxy‐4‐[5‐(penta‐1,3‐diyn‐1‐yl)thiophen‐2‐yl]but‐3‐yn‐2‐yl acetate ( 7 ), 4‐(2,2′‐bithiophen‐5‐yl)but‐3‐yne‐1,2‐diol ( 8 ), and 4‐[5‐(penta‐1,3‐diyn‐1‐yl)thiophen‐2‐yl]but‐3‐yne‐1,2‐diol ( 9 ), isolated from the roots of Echinops transiliensis, were studied as larvicides against Aedes aegypti. Structural differences among compounds 3, 5 , and 8 consisted in differing AcO and OH groups attached to C(3″) and C(4″), and resulted in variations in efficacy. Terthiophene 1 showed the highest activity (LC50, 0.16 μg/ml) among compounds 1 – 9 , followed by bithiophene compounds 3 (LC50, 4.22 μg/ml), 5 (LC50, 7.45 μg/ml), and 8 (LC50, 9.89 μg/ml), and monothiophene compounds 9 (LC50, 12.45 μg/ml), 2 (LC50, 14.71 μg/ml), 4 (LC50, 17.95 μg/ml), 6 (LC50, 18.55 μg/ml), and 7 (LC50, 19.97 μg/ml). These data indicated that A. aegypti larvicidal activities of thiophenes increase with increasing number of thiophene rings, and the most important active site in the structure of thiophenes could be the tetrahydro‐thiophene moiety. In bithiophenes, 3, 5 , and 8 , A. aegypti larvicidal activity increased with increasing number of AcO groups attached to C(3″) or C(4″), indicating that AcO groups may play an important role in the larvicidal activity.  相似文献   

8.
Introduction – The aerial part Eupatorium lindleyanum is commonly used as an antipyretic and detoxicant clinically in traditional Chinese medicine. Our previous research showed that germacrane sesquiterpene lactones were its main active constituents, so the development of rapid and accurate methods for the identification of the sesquiterpene lactones is of great significance. Objective – To develop an HPLC‐PDA‐ESI‐MS/MS method capable for simple and rapid analysis of germacrane sesquiterpene lactones in the aerial part E. lindleyanum. Methodology – High‐performance liquid chromatography‐photodiode array detection‐electrospray ionization‐tandem mass spectrometry was used to analyze germacrane sesquiterpene lactones of Eupatorium lindleyanum. The fragmentation behavior of germacrane sesquiterpene lactones in a Micromass Q/TOF Mass Spectrometer was discussed, and 9 germacrane sesquiterpene lactones were identified by comparison of their characteristic data of HPLC and MS analyses with those obtained from reference compounds. Results – The investigated germacrane sesquiterpene lactones were identified as eupalinolides C (1), 3β‐acetoxy‐8β‐(4′‐hydroxy‐tigloyloxy)‐14‐hydroxy‐costunolide (2), eupalinolides A (3), eupalinolides B (4), eupalinolides E (5), 3β‐acetoxy‐8β‐(4′‐oxo‐tigloyloxy)‐14‐hydroxy‐heliangolide (6), 3β‐acetoxy‐8β‐(4′‐oxo‐ tigloyloxy)‐14‐hydroxy‐costunolide (7), hiyodorilactone B (8), and 3β‐acetoxy‐8β‐(4′‐hydroxy‐tigloyloxy)‐ costunolide (9). Compounds 6, 7 and 9 were reported for the first time. Conclusion – HPLC‐PDA‐ESI‐MS/MS provides a new powerful approach to identify germacrane sesquiterpene lactones in E. lindleyanum rapidly and accurately. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

9.
Three new iridoids named as pediverticilatasin A – C ( 1 – 3 , resp.), together with five known iridoids ( 4 – 8 , resp.) were isolated from the whole plants of Pedicularis verticillata. The structures of three new compounds were identified as (1S,7R)‐1‐ethoxy‐1,5,6,7‐tetrahydro‐7‐hydroxy‐7‐methylcyclopenta[c]pyran‐4(3H)‐one ( 1 ), (1S,4aS,7R,7aS)‐1‐ethoxy‐1,4a,5,6,7,7a‐hexahydro‐7‐hydroxy‐7‐methylcyclopenta[c]pyran‐4‐carboxylic acid ( 2 ), (1S,4aS,7R,7aS)‐1‐ethoxy‐1,4a,5,6,7,7a‐hexahydro‐7‐hydroxy‐7‐methylcyclopenta[c]pyran‐4‐carbaldehyde ( 3 ). Their structures were elucidated on the basis of spectroscopic methods and compared with the NMR spectra data in the literature. All compounds were evaluated for their anti‐complementary activity on the classical pathway of the complement system in vitro. Among which, compounds 1 , 3 , and 6 exhibited anti‐complementary effects with CH50 values ranging from 0.43 to 1.72 mm , which are plausible candidates for developing potent anti‐complementary agents.  相似文献   

10.
Two new flavones, 6,7‐methylenedioxy‐4‐hydroxypeltogynan‐7′‐one ( 1 ), cochliophilin B ( 2 ), as well as two known ones, cochliophilin A ( 3 ) and 6‐methoxy‐7‐hydroxy flavone ( 4 ), were isolated from the ethanol extract of the root of Phytolacca acinosa Roxb . Compound 1 is a flavanol framework with one δ‐lactone unit, which is rather rare in nature. The structures of the new compounds were determined on the basis of extensive spectroscopic (IR, MS, 1D‐ and 2D‐NMR) analyses, the absolute configuration of 1 was established by comparing experimental and calculated electronic circular dichroism spectra. The structures of known compounds were fixed by comparison with literatures data. Compounds 2 and 4 showed modest inhibitory activities against BEL‐7402 cell line, with IC50 values of 28.22 and 39.16 μmol/L, respectively.  相似文献   

11.
The reduction of the axially chiral N‐(o‐aryl)‐5,5‐dimethyl‐2,4‐oxazolidinediones by NaBH4 yielded axially chiral N‐(o‐aryl)‐4‐hydroxy‐5,5‐dimethyl‐2‐oxazolidinone enantiomers having a chiral center at C‐4, with 100% diastereoselectivity as has been shown by their 1H and 13C NMR spectra and by enantioselective HPLC analysis. The resolved enantiomeric isomers were found to interconvert thermally through an aldehyde intermediate formed upon ring cleavage via a latent ring‐chain‐ring tautomerization. It was found that the rate of enantiomerization depended on the size and the electronic effect of the ortho substituent present on the aryl ring bonded to the nitrogen of the heterocycle. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

12.
A series of novel indenopyrazole derivatives 2a‐j and 3a‐j were synthesized from the reaction of 1‐(4‐(hydroxy(1‐oxo‐1,3‐dihydro‐2 H‐inden‐2‐ylidene)methyl)phenyl)‐3‐phenylurea derivatives 1a‐j with hydrazine and phenylhydrazine, respectively. The obtained novel indenopyrazoles ( 2a‐j and 3a‐j ) were evaluated for anticancer activity against HeLa and C6 cell lines. Antiproliferative activity was determined by the BrdU proliferation ELISA assay; 2a , 2b , 2d , 2h , and 3h were found to be the most active compounds. The compounds were also screened for antimicrobial activity, and all compounds showed moderate activity against used microorganisms.  相似文献   

13.
Five new geminal aminocycloalkanephosphonic acids ( 4 – 8 ) containing both an aromatic ring and a cycloalkane ring were synthesized and evaluated as potential inhibitors of buckwheat phenylalanine ammonia‐lyase (PAL). Within the set of compounds which are related to 2‐aminoindane‐2‐phosphonic acid (AIP, 3 ), a known powerful inhibitor of PAL, racemic 1‐aminobenzocyclobutene‐1‐phosphonic acid ( 4 ), was six times weaker than AIP as an in vitro inhibitor of buckwheat PAL, but six times stronger than AIP as an in vivo inhibitor of phenylalanine‐derived anthocyanin synthesis in buckwheat.  相似文献   

14.
Condensation of 1,3‐dihydro‐2,3‐dioxo‐2H‐indoles (1a–c) with galactaric acid bis hydrazide (2) gave the corresponding galactaric acid bis[2‐(1,2‐dihydro‐2‐oxo‐3H‐indol‐3‐ylidene)hydrazides] (3a–c). Acetylation of the latter compounds with acetic anhydride in the presence of pyridine at ambient temperature gave the 2,3,4,5‐tetra‐O‐acetylgalactaric acid bis[2‐(1,2‐dihydro‐2‐oxo‐1‐substituted‐3H‐indol‐3‐ylidene)hydrazides] (4b–d). Heterocyclization of the tetra‐O‐acetates 4b–d by heating with thionyl chloride afforded the double headed acyclo C‐nucleosides: 1,2,3,4‐tetra‐O‐acetyl‐1,4‐bis{9‐substituted‐1,3,4‐oxadiazino[6,5‐b]indol‐2‐yl‐1‐ium}‐galacto‐tetritol dichlorides (5b–d). Structures of the prepared compounds were elucidated from their spectral properties.  相似文献   

15.
The in vivo sex‐linked recessive lethal test was carried out in Drosophila melanogaster to investigate whether or not five substituted 4‐hydroxy‐2H‐chromen‐2‐ones can modulate the genotoxicity of the well‐established mutagenic agent ethyl methanesulfonate (EMS). For this purpose, 3 days old Canton S males were treated with the potent mutagen EMS alone in concentration of 0.75 ppm, as well as in combination with one of the five 4‐hydroxycoumarins, namely diethyl 2‐(1‐(4‐hydroxy‐2‐oxo‐2H‐chromen‐3‐yl)ethylidene)malonate ( 2b ), 3‐(1‐(4‐hydroxy‐2‐oxo‐2H‐chromen‐3‐yl)ethylidene)pentane‐2,4‐dione ( 6b ), 4‐(4‐(4‐hydroxy‐2‐oxo‐2H‐chromen‐3‐yl)thiazol‐2‐ylamino) benzenesulfonic acid ( 4c ), 4‐hydroxy‐3‐(2‐(2‐nitropheny lamino)thiazol‐4‐yl)‐2H‐chromen‐2‐one ( 9c ), and (E)‐4‐hydroxy‐3‐(1‐(m‐tolylimino)ethyl)‐2H‐chromen‐2‐one ( 5d ), in concentration of 70 ppm. The frequency of germinative mutations increased significantly after the treatment with EMS and decreased after treatments with coumarins. The maximum reduction was observed after treatments with 2b , 6b , 4c , and 5d . By the formation of hydrogen bonds or electrostatic interactions with O6 of DNA guanine, tested coumarins prevent EMS‐induced alkylation. The results indicate a protective role of five 4‐hydroxycoumarins under the action of a strong mutagen. © 2012 Wiley Periodicals, Inc. J Biochem Mol Toxicol 26:322–330, 2012; View this article online at wileyonlinelibrary.com . DOI 10.1002/jbt.21426  相似文献   

16.
A pathway for the biotransformation of the environmental pollutant and high‐molecular weight polycyclic aromatic hydrocarbon (PAH) benzo[k]fluoranthene by a soil bacterium was constructed through analyses of results from liquid chromatography negative electrospray ionization tandem mass spectrometry (LC/ESI(–)‐MS/MS). Exposure of Sphingobium sp. strain KK22 to benzo[k]fluoranthene resulted in transformation to four‐, three‐ and two‐aromatic ring products. The structurally similar four‐ and three‐ring non‐alternant PAHs fluoranthene and acenaphthylene were also biotransformed by strain KK22, and LC/ESI(–)‐MS/MS analyses of these products confirmed the lower biotransformation pathway proposed for benzo[k]fluoranthene. In all, seven products from benzo[k]fluoranthene and seven products from fluoranthene were revealed and included previously unreported products from both PAHs. Benzo[k]fluoranthene biotransformation proceeded through ortho‐cleavage of 8,9‐dihydroxy‐benzo[k]fluoranthene to 8‐carboxyfluoranthenyl‐9‐propenic acid and 9‐hydroxy‐fluoranthene‐8‐carboxylic acid, and was followed by meta‐cleavage to produce 3‐(2‐formylacenaphthylen‐1‐yl)‐2‐hydroxy‐prop‐2‐enoic acid. The fluoranthene pathway converged with the benzo[k]fluoranthene pathway through detection of the three‐ring product, 2‐formylacenaphthylene‐1‐carboxylic acid. Production of key downstream metabolites, 1,8‐naphthalic anhydride and 1‐naphthoic acid from benzo[k]fluoranthene, fluoranthene and acenaphthylene biotransformations provided evidence for a common pathway by strain KK22 for all three PAHs through acenaphthoquinone. Quantitative analysis of benzo[k]fluoranthene biotransformation by strain KK22 confirmed biodegradation. This is the first pathway proposed for the biotransformation of benzo[k]fluoranthene by a bacterium.  相似文献   

17.
Papulaspora immersa H. H. Hotson was isolated from roots and leaves of Smallanthus sonchifolius (Poepp. and Endl. ) H. Rob. (Asteraceae), traditionally known as Yacon. The fungus was cultured in rice, and, from the AcOEt fraction, 14 compounds were isolated. Among them, (22E,24R)‐8,14‐epoxyergosta‐4,22‐diene‐3,6‐dione ( 4 ), 2,3‐epoxy‐1,2,3,4‐tetrahydronaphthalene‐c‐1,c‐4,8‐triol ( 10 ), and the chromone papulasporin ( 13 ) were new secondary metabolites. The spectral data of the known natural products were compared with the literature data, and their structures were established as the (24R)‐stigmast‐4‐en‐3‐one ( 1 ), 24‐methylenecycloartan‐3β‐ol ( 2 ), (22E,24R)‐ergosta‐4,6,8(14),22‐tetraen‐3‐one ( 3 ), (?)‐(3R,4R)‐4‐hydroxymellein ( 5 ), (?)‐(3R)‐5‐hydroxymellein ( 6 ), 6,8‐dihydroxy‐3‐methylisocoumarin ( 7 ), (?)‐(4S)‐4,8‐dihydroxy‐α‐tetralone ( 8 ), naphthalene‐1,8‐diol ( 9 ), 6,7,8‐trihydroxy‐3‐methylisocoumarin ( 11 ), 7‐hydroxy‐2,5‐dimethylchromone ( 12 ), and tyrosol ( 14 ). Compound 4 showed the highest cytotoxic activity against the human tumor cell lines MDA‐MB435 (melanoma), HCT‐8 (colon), SF295 (glioblastoma), and HL‐60 (promyelocytic leukemia), with IC50 values of 3.3, 14.7, 5.0 and 1.6 μM , respectively. Strong synergistic effects were also observed with compound 5 and some of the isolated steroidal compounds.  相似文献   

18.
Enantiopure 3((R)‐ and 3((S)‐1‐phenylethyl)‐4‐oxazoline‐2‐ones were evaluated as chiral building blocks for the divergent construction of heterocycles with stereogenic quaternary centers. The N‐(R)‐ or N‐(S)‐1‐phenylethyl group of these compounds proved to be an efficient chiral auxiliary for the asymmetric induction of the 4‐ and 5‐positions of the 4‐oxazolin‐2‐one ring through thermal and MW‐promoted nucleophilic conjugated addition to Michael acceptors and alkyl halides. The resulting adducts were transformed via a cascade process into fused six‐membered carbo‐ and heterocycles. The structure of the reaction products depended on the electrophiles and reaction conditions used. Alternative isomeric 4‐methylene‐2‐oxazolidinones served as chiral precursors for a versatile and divergent approach to highly substituted cyclic carbamates. DFT quantum calculations showed that the formation of bicyclic pyranyl compounds was generated by a diastereoselective concerted hetero‐Diels‐Alder cycloaddition.  相似文献   

19.
《Luminescence》2002,17(5):305-312
Four bicyclic dioxetanes bearing a phenolic substituent, 3‐tert‐butyldimethylsiloxy‐4‐chlorophenyl ( 3a ), 5‐tert‐butyldimethylsiloxy‐4‐chloro‐2‐ethylphenyl ( 3b ), 5‐tert‐butyldimethylsiloxy‐2‐ethylphenyl ( 3c ), and 3‐tert‐butyldimethylsiloxy‐4‐ethylphenyl ( 3d ), were synthesized. All dioxetanes 3a – 3d gave intense blue light on treatment with tetrabutylammonium fluoride (TBAF) in DMSO or acetonitrile. Kinetic study on the fluoride‐induced CIEEL decay of these dioxetanes 3a – 3d and the parent dioxetane 2b revealed that the para‐substitution with chlorine on the phenolic moiety of dioxetane increases free energy of activation (ΔG?), while the para‐substitution with ethyl on the aryl decreases ΔG?. On the other hand, substitution with an ethyl at the ortho‐position instead of the para‐position was found to increase ΔG? and to suppress the CIEEL decay. This fact is attributed to the steric factor of the ortho‐ethyl group which would prevent the aromatic ring from rotating freely around the axis joined to the peroxide ring, and supports the suggestion for a CIEEL‐active dioxetane bearing a phenolic moiety that an intramolecular electron transfer occurs preferentially from the phenolic donor to O–O of the dioxetane ring, when the aromatic ring lies in a certain conformation(s). Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

20.
RS‐4‐(4‐Hydroxyphenyl)‐2‐butanol (rhododendrol, RD) was used as a skin‐whitening agent until it was reported to induce leukoderma in July 2013. To explore the mechanism underlying its melanocyte toxicity, we characterized the tyrosinase‐catalyzed oxidation of RD using spectrophotometry and HPLC. Oxidation of RD with mushroom tyrosinase rapidly produced RD‐quinone, which was quickly converted to 2‐methylchromane‐6,7‐dione (RD‐cyclic quinone) and RD‐hydroxy‐p‐quinone through cyclization and addition of water molecule, respectively. RD‐quinone and RD‐cyclic quinone were identified as RD‐catechol and RD‐cyclic catechol after NaBH4 reduction. Autoxidation of RD‐cyclic catechol produced superoxide radical. RD‐quinone and RD‐cyclic quinone quantitatively bound to thiols such as cysteine and GSH. These results suggest that the melanocyte toxicity of RD is caused by its tyrosinase‐catalyzed oxidation through production of RD‐cyclic quinone which depletes cytosolic GSH and then binds to essential cellular proteins through their sulfhydryl groups. The production of ROS through autoxidation of RD‐cyclic catechol may augment the toxicity.  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