首页 | 本学科首页   官方微博 | 高级检索  
文章检索
  按 检索   检索词:      
出版年份:   被引次数:   他引次数: 提示:输入*表示无穷大
  收费全文   184篇
  免费   0篇
  国内免费   1篇
  2020年   1篇
  2019年   4篇
  2018年   3篇
  2017年   4篇
  2016年   2篇
  2015年   4篇
  2014年   4篇
  2013年   12篇
  2012年   2篇
  2011年   12篇
  2010年   1篇
  2009年   9篇
  2008年   15篇
  2007年   11篇
  2006年   5篇
  2005年   6篇
  2004年   14篇
  2003年   10篇
  2002年   7篇
  2001年   1篇
  2000年   4篇
  1999年   1篇
  1998年   4篇
  1997年   3篇
  1996年   2篇
  1995年   2篇
  1993年   3篇
  1991年   1篇
  1990年   3篇
  1989年   3篇
  1987年   2篇
  1986年   2篇
  1984年   6篇
  1983年   5篇
  1982年   4篇
  1981年   5篇
  1980年   3篇
  1979年   4篇
  1971年   1篇
排序方式: 共有185条查询结果,搜索用时 31 毫秒
111.
The efficacy of the diphenol curcumin as a cancer chemopreventive agent is limited by its chemical and metabolic instability. Non-enzymatic degradation has been described to yield vanillin, ferulic acid, and feruloylmethane through cleavage of the heptadienone chain connecting the phenolic rings. Here we provide evidence for an alternative mechanism, resulting in autoxidative cyclization of the heptadienone moiety as a major pathway of degradation. Autoxidative transformation of curcumin was pH-dependent with the highest rate at pH 8 (2.2 μM/min) and associated with stoichiometric uptake of O(2). Oxidation was also catalyzed by recombinant cyclooxygenase-2 (COX-2) (50 nm; 7.5 μM/min), and the rate was increased ≈10-fold by the addition of 300 μM H(2)O(2). The COX-2 catalyzed transformation was inhibited by acetaminophen but not indomethacin, suggesting catalysis occurred by the peroxidase activity. We propose a mechanism of enzymatic or autoxidative hydrogen abstraction from a phenolic hydroxyl to give a quinone methide and a delocalized radical in the heptadienone chain that undergoes 5-exo cyclization and oxygenation. Hydration of the quinone methide (measured by the incorporation of O-18 from H(2)(18)O) and rearrangement under loss of water gives the final dioxygenated bicyclopentadione product. When curcumin was added to RAW264.7 cells, the bicyclopentadione was increased 1.8-fold in cells activated by LPS; vanillin and other putative cleavage products were negligible. Oxidation to a reactive quinone methide is the mechanistic basis of many phenolic anti-cancer drugs. It is possible, therefore, that oxidative transformation of curcumin, a prominent but previously unrecognized reaction, contributes to its cancer chemopreventive activity.  相似文献   
112.
《Free radical research》2013,47(4-6):391-399
The exposure of mammalian cells to toxic concentrations of redox cycling and alkylating quinones causes marked changes in cell surface structure knoun as plasma membrane blebbing. These alterations are associated with the redistribution of plasma membrane proteins and the disruption of the normal organization of the cytoskeletal microfilaments which appears 10 be due mainly to actin cross-linking and dissociation of α-actinin from the actin network. The major hiochemical mechanisms responsible for these effects seem to involve the depletion of cytoskeletal protein dfhydryl groups and the increase cytosolic Ca2+ concentration following the alkylation/oxidation of free sulfhydryl groups in several Ca2+ transport systems. Depletion of intracellular ATP is also associated with quinone-induced plasma menibrane blebbing. However, ATP depletion occurs well after the onset of the morphological changes. and thus it does not seem to be causatively related to their appeardncc Thiol reductants. such as dithiothreitol. efficiently prevent the oxidation of cytoskeletal protein thiols. the increase in cytosolic free Ca2+ concentration and cell blebbing induced by redox cycling. but not alkylating. quinones. These results demonstrate that alkylating and redox cycling quinones cause siinilar structural and biochemical modifications of the cytoskeleton by means of different mechanisms. namely alkylation and oxidation of critical sulfhydryl groups  相似文献   
113.
Benzofuranic analogues of MCA-NAT (5-methoxycarbonylamino-N-acetyltryptamine) have been synthesized and evaluated as melatonin receptor ligands. Introduction of a methoxycarbonylamino substituent in the C-5 position of the benzofurane nucleus obtains MT3 selective ligands. This selectivity can be modulated with suitable variations of the C-5 position and the acyl group on the C-3 side chain.  相似文献   
114.
This review will focus on how X-ray crystallographic studies of copper-containing amine oxidases have complemented the solution, kinetic, and spectroscopic research on this ubiquitous class of enzymes. These enzymes not only contain a copper ion at the active site, but also a unique organic cofactor, 2,4,5-trihydroxyphenylalanine quinone (TPQ), which is absolutely required for catalysis. Structural data have not only shed light on the catalytic mechanism of the enzyme, which converts primary amines, using molecular oxygen, to aldehydes, ammonia, and hydrogen peroxide, but also on biogenesis of the cofactor. The cofactor is derived from a tyrosine in the enzyme amino acid sequence and requires only the addition of copper(II) and molecular oxygen in a self-processing event.  相似文献   
115.
Soluble quinoprotein dehydrogenases oxidize a wide range of sugar, alcohol, amine, and aldehyde substrates. The physiological electron acceptors for these enzymes are not pyridine nucleotides but are other soluble redox proteins. This makes these enzymes and their electron acceptors excellent systems with which to study mechanisms of long-range interprotein electron transfer reactions. The tryptophan tryptophylquinone (TTQ)-dependent methylamine dehydrogenase (MADH) transfers electrons to a blue copper protein, amicyanin. It has been possible to alter the rate of electron transfer by using different redox forms of MADH, varying reaction conditions, and performing site-directed mutagenesis on these proteins. From kinetic and thermodynamic analyses of the reaction rates, it was possible to determine whether a change in rate is due a change in Delta G(0), electronic coupling, reorganization energy or kinetic mechanism. Examples of each of these cases are discussed in the context of the known crystal structures of the electron transfer protein complexes. The pyrroloquinoline quinone (PQQ)-dependent methanol dehydrogenase transfers electrons to a c-type cytochrome. Kinetic and thermodynamic analyses of this reaction indicated that this electron transfer reaction was conformationally coupled. Quinohemoproteins possess a quinone cofactor as well as one or more c-type hemes within the same protein. The structures of a PQQ-dependent quinohemoprotein alcohol dehydrogenase and a TTQ-dependent quinohemoprotein amine dehydrogenase are described with respect to their roles in intramolecular and intermolecular protein electron transfer reactions.  相似文献   
116.
Cytotoxicity of 1,4-naphthoquinones has been attributed to intracellular reactive oxygen species (ROS) generation through one-electron-reductase-mediated redox cycling and to arylation of cellular nucleophiles. Here, however, we report that in a subclone of lung epithelial A549 cells (A549-S previously called A549-G4S (Watanabe, et al., Am. J. Physiol. 283 (2002) L726-736), the mechanism of ROS generation by menadione and by 2,3-dimethoxy-1,4-naphthoquinone (DMNQ), and therefore that of cytotoxicity, differs from the paradigm. Ninety percent of H(2)O(2) generation by both the quinones can be prevented by dicumarol, an inhibitor of NAD(P)H quinone oxidoreductase (NQO1), at the submicromolar level, regardless of the quinone concentrations. Exogenous SOD also inhibits H(2)O(2) production at low but not high concentrations of the quinones, especially DMNQ. Thus, at low quinone concentrations, superoxide-driven hydroquinone autoxidation accounts for more than half of H(2)O(2) generation by both quinones, whereas at high quinone concentrations, especially for DMNQ, comproportionation-driven hydroquinone autoxidation becomes the predominant mechanism. Hydroquinone autoxidation appears to occur predominantly in the extracellular environment than in the cytosol as extracellular catalase can dramatically attenuate quinone-induced cytotoxicity throughout the range of quinone concentrations, whereas complete inactivation of endogenous catalase or complete depletion of intracellular glutathione has only a marginal effect on their cytotoxicity. Finally, we show evidence that ROS production is a consequence of the compensatory defensive role of NQO1 against quinone arylation.  相似文献   
117.
The anaerobically expressed fumarate reductase and aerobically expressed succinate dehydrogenase from Escherichia coli comprise two different classes of succinate:quinone oxidoreductases (SQR), often termed respiratory complex II. The X-ray structures of both membrane-bound complexes have revealed that while the catalytic/soluble domains are structurally similar the quinone binding domains of the enzyme complexes are significantly different. These results suggest that the anaerobic and aerobic forms of complex II have evolved different mechanisms for electron and proton transfer in their respective membrane domains.  相似文献   
118.
Coussaric acid (1), a triterpenoid based on an ursane skeleton, and an oleanane-type triterpene acid, 3-epi-spathodic acid (2), as well as four known compounds, barbinervic acid, scutellaric acid, stigmasterol and stigmasterol glucoside, have been isolated from an EtOAc-soluble extract of the stems of Coussarea brevicaulis. The structures of compounds 1 and 2 were elucidated on the basis of spectroscopic investigation, and single-crystal X-ray crystallography was used to confirm the structure of 1. The absolute stereochemistry of 1 was established by chemical transformations and by the Mosher ester procedure. The potential of the isolates and chemical transformation products to induce quinone reductase was evaluated in mouse Hepa lclc7 hepatoma cells.  相似文献   
119.
Quinone reductase 2 is a mammalian cytosolic FAD-dependent enzyme, the activity of which is not supported by conventional nicotinamide nucleotides. An endobiotic substrate has never been reported for this enzyme nor a set of molecular tools, such as inhibitors. In the present work, we used the recombinant human enzyme, expressed in CHO cells for the systematic screening of both co-substrates and substrates. The co-substrates survey showed that the natural occurring compound, N-ribosylnicotinamide, was a poor co-substrate. The synthetic N-benzylnicotinamide is a better one compared to any other compounds tested. We found that tetrahydrofolic acid acted as a co-substrate for the reduction of menadione catalysed by quinone reductase 2, although with poor potency (Km approximately 2 mM). Among a series of commercially available quinones, a single one was found to be substrate of quinone reductase 2, in the presence of N-benzyldihydronicotinamide: coenzyme Q0. Finally, we tested a series of 197 flavonoids as potential inhibitors. We found apigenin, genistein or kaempferol as good inhibitor of quinone reductase 2 activity with IC50 in the 100 nM range. These compounds, co-substrate, substrate and inhibitors will permit to better know this enzyme, the role of which is still poorly understood.  相似文献   
120.
Phenotypic and phylogenetic studies were performed on unidentified Gram-negative staining, haloalkaliphilic aerobe and protease producer Salinivibrio-like organism recovered from a saltish spring with algal mat in the “Pozzo del Sale” site (Salt's Well) in the Campania Region (South Italy). Phylogenetic analysis based on comparison of 16S rRNA gene sequences demonstrated that the isolate was related to species of Salinivibrio genus. The DNA–DNA hybridization of the type strain 18AGT with the most related Salinivibrio costicola subsp. costicola showed a re-association value of 72%. Based on the phenotypic distinctiveness of 18AGT strain and molecular, chemical and genetic evidence, it is proposed that strain 18AGT can be classified as S. costicola subsp. alcaliphilus, subsp. nov. The type strain of S. costicola subsp. alcaliphilus, is ATCC BAA-952T; DSM 16359T.  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