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   373篇
  免费   33篇
  国内免费   1篇
  2022年   3篇
  2021年   8篇
  2019年   4篇
  2018年   3篇
  2017年   9篇
  2016年   11篇
  2015年   8篇
  2014年   21篇
  2013年   14篇
  2012年   22篇
  2011年   27篇
  2010年   17篇
  2009年   16篇
  2008年   15篇
  2007年   13篇
  2006年   19篇
  2005年   15篇
  2004年   11篇
  2003年   13篇
  2002年   10篇
  2001年   10篇
  2000年   6篇
  1999年   5篇
  1998年   13篇
  1997年   8篇
  1996年   5篇
  1995年   3篇
  1994年   5篇
  1993年   7篇
  1992年   3篇
  1991年   7篇
  1990年   9篇
  1989年   7篇
  1988年   5篇
  1987年   7篇
  1986年   7篇
  1985年   4篇
  1983年   4篇
  1982年   3篇
  1981年   3篇
  1979年   7篇
  1978年   3篇
  1977年   2篇
  1976年   3篇
  1975年   3篇
  1971年   1篇
  1970年   1篇
  1969年   2篇
  1968年   2篇
  1959年   1篇
排序方式: 共有407条查询结果,搜索用时 15 毫秒
81.
The transfer of electrons and protons between membrane-bound respiratory complexes is facilitated by lipid-soluble redox-active quinone molecules (Q). This work presents a structural analysis of the quinone-binding site (Q-site) identified in succinate:ubiquinone oxidoreductase (SQR) from Escherichia coli. SQR, often referred to as Complex II or succinate dehydrogenase, is a functional member of the Krebs cycle and the aerobic respiratory chain and couples the oxidation of succinate to fumarate with the reduction of quinone to quinol (QH(2)). The interaction between ubiquinone and the Q-site of the protein appears to be mediated solely by hydrogen bonding between the O1 carbonyl group of the quinone and the side chain of a conserved tyrosine residue. In this work, SQR was co-crystallized with the ubiquinone binding-site inhibitor Atpenin A5 (AA5) to confirm the binding position of the inhibitor and reveal additional structural details of the Q-site. The electron density for AA5 was located within the same hydrophobic pocket as ubiquinone at, however, a different position within the pocket. AA5 was bound deeper into the site prompting further assessment using protein-ligand docking experiments in silico. The initial interpretation of the Q-site was re-evaluated in the light of the new SQR-AA5 structure and protein-ligand docking data. Two binding positions, the Q(1)-site and Q(2)-site, are proposed for the E. coli SQR quinone-binding site to explain these data. At the Q(2)-site, the side chains of a serine and histidine residue are suitably positioned to provide hydrogen bonding partners to the O4 carbonyl and methoxy groups of ubiquinone, respectively. This allows us to propose a mechanism for the reduction of ubiquinone during the catalytic turnover of the enzyme.  相似文献   
82.
The tocopherol transfer protein (TTP) is a member of the CRAL-TRIO family of lipid binding proteins that facilitates vitamin E transfer between membrane vesicles in vitro. In cultured hepatocytes, TTP enhances the secretion of tocopherol to the media; presumably, tocopherol transfer is at the basis of this biological activity. The mechanism underlying ligand transfer by TTP is presently unknown, and available tools for monitoring this activity suffer from complicated assay procedure and poor sensitivity. We report the characterization of a fluorescent vitamin E analogue, (R)-2,5,7,8-tetramethylchroman-2-[9-(7-nitrobenz[1,2,5]oxadiazol-4-ylamino)nonyl]chroman-6-ol (NBD-TOH), as a sensitive and convenient probe for the ligand binding and transfer activities of TTP. Upon binding to TTP, NBD-TOH fluorescence is blue shifted, and its intensity is greatly enhanced. We used these properties to accurately determine the affinity of NBD-TOH to TTP. The analogue binds to TTP reversibly and with high affinity (K(d) = 8.5 +/- 6 nM). We determined the affinity of NBD-TOH to a TTP protein in which lysine 59 is replaced with a tryptophan. When occurring in humans, this heritable mutation causes the ataxia with vitamin E deficiency (AVED) disorder. We find that the affinity of NBD-TOH to this mutant TTP is greatly diminished (K(d) = 71 +/- 19 nM). NBD-TOH functioned as a sensitive fluorophore in fluorescent resonance energy transfer (FRET) experiments. Using the fluorescent lipids TRITC-DHPE or Marina Blue-DHPE as a donor or an acceptor for NBD-TOH fluorescence, we obtained high-resolution kinetic data for tocopherol movement out of lipid bilayers, a key step in the TTP-facilitated ligand transfer reaction.  相似文献   
83.
Three new structures of Escherichia coli succinate-quinone oxidoreductase (SQR) have been solved. One with the specific quinone-binding site (Q-site) inhibitor carboxin present has been solved at 2.4 Å resolution and reveals how carboxin inhibits the Q-site. The other new structures are with the Q-site inhibitor pentachlorophenol and with an empty Q-site. These structures reveal important details unresolved in earlier structures. Comparison of the new SQR structures shows how subtle rearrangements of the quinone-binding site accommodate the different inhibitors. The position of conserved water molecules near the quinone binding pocket leads to a reassessment of possible water-mediated proton uptake networks that complete reduction of ubiquinone. The dicarboxylate-binding site in the soluble domain of SQR is highly similar to that seen in high resolution structures of avian SQR (PDB 2H88) and soluble flavocytochrome c (PDB 1QJD) showing mechanistically significant structural features conserved across prokaryotic and eukaryotic SQRs.Succinate:quinone oxidoreductase (SQR,4 succinate dehydrogenase) and menaquinol:fumarate oxidoreductase (QFR, fumarate reductase), members of the Complex II family, are homologous integral membrane proteins which couple the interconversion of succinate and fumarate with quinone and quinol (14). SQR is a key enzyme in the Krebs cycle, oxidizing succinate to fumarate during aerobic growth and reducing quinone to quinol and, thus, acts as a direct link between the Krebs cycle and the respiratory chain. QFR is found in anaerobic or facultative bacteria and lower eukaryotes, where it couples the oxidation of reduced quinones to the reduction of fumarate (1, 4). Escherichia coli SQR has four subunits, two hydrophilic subunits exposed to the cytoplasm (SdhA and SdhB), which interact with two hydrophobic membrane-intrinsic subunits (SdhC and SdhD) (5). SdhA contains the dicarboxylate-binding site and a covalently bound FAD cofactor which cycles between FAD and FADH2 redox states during succinate oxidation (6). The electrons from succinate oxidation are sequentially transferred via a [2Fe-2S], a [4Fe-4S], and a [3Fe-4S] iron-sulfur cluster relay system in SdhB to a quinone-binding site (QP) located at the interface of the SdhB, SdhC, and SdhD subunits. SdhC and SdhD are both composed of three transmembrane helices and coordinate a low spin b-type heme via His residues contributed by each subunit (7, 8).The first structural information about members of the Complex II family came from x-ray structures of the QFR enzymes from E. coli at 3.3 Å resolution (9) and Wolinella succinogenes at 2.2 Å resolution (10). These structures revealed details of the overall architecture of the subunits, the position of key redox cofactors, the electron transfer pathway, and the quinone-binding sites. At around the same time, the structures of soluble fumarate reductases found in anaerobic and microaerophilic bacteria and structurally homologous to the flavoprotein subunit of Complex II were solved by x-ray crystallography (1). Analysis of these soluble fumarate reductases has proven particularly informative in describing the mechanism of fumarate reduction and succinate oxidation at the dicarboxylate-binding site (1114).Structures of SQRs lagged behind those of the QFRs until the structure of the E. coli enzyme was solved at 2.6 Å (15). This structure, solved in space group R32, revealed that the E. coli enzyme is packed as a trimer. The structures of the SdhA and SdhB subunits were highly similar to those of E. coli and W. succinogenes QFRs, but the transmembrane SdhC and SdhD subunits showed differences compared with their QFR counterparts. The structure revealed the position of the redox sites and the dicarboxylate- and quinone-binding (Q) sites. The heme b molecule was shown to lie away from the electron transfer pathway, suggesting electrons are preferentially transferred from the [3Fe-4S] cluster to ubiquinone, on the grounds of the edge-to-edge distances and redox potentials of the relevant groups. The structure revealed density in the Q-site that was interpreted as ubiquinone, and the position of the binding site was confirmed by the structure of the E. coli enzyme co-crystallized with the Q-site inhibitor 2-(1-methyl-hexyl)-4,6-dinitrophenol (DNP-17, PDB code 1NEN (15)). The E. coli enzyme was subsequently co-crystallized with the Q-site inhibitor Atpenin A5 (AA5) (PDB code 2ACZ (16)). This inhibitor was bound deeper into the quinone-binding site than ubiquinone or DNP-17, suggesting that there are two binding positions for ubiquinone in its binding site. The structure also identified a water-mediated proton pathway, proposed to deliver protons to the quinone-binding site. The first structure of a mitochondrial SQR was from porcine heart at 2.4 Å resolution (PDB code 1ZOY (17). This structure revealed a monomer in the asymmetric unit, suggesting that mitochondrial SQRs were likely to function as monomers. Superposition of the porcine and E. coli SQR structures revealed the high structural similarity of the SdhA and SdhB subunits and the conservation in position of the redox cofactors. Larger divergences were observed in the transmembrane subunits.Further structural information about SQRs was obtained by analysis of structures of avian SQR crystallized with oxaloacetate (2.2 Å resolution, PDB code 1YQ3), with 3-nitropropionate (2.4 Å resolution, PDB code 1YQ4), and with the Q-site inhibitor carboxin (2.1 Å resolution, PDB code 2FBW) (18). These structures revealed important differences in the position of key residues in the dicarboxylate-binding site compared with the E. coli and porcine structures. Arg-297 (equivalent to Arg-298 in porcine and Arg-286 in E. coli SQRs) was ideally located to act as a general base catalyst, accepting a proton during dehydrogenation of succinate, as in the soluble Shewanella flavocytochrome c3 (PDB code 1QJD) (11), suggesting conservation of mechanism between these distantly related enzymes. An unusual cis-serine peptide bond was proposed to position another arginine residue for binding dicarboxylates. Density for the dicarboxylate in 1YQ3 and 2FBW was shown to be distinctly non-planar and could be modeled by the “malate-like intermediate” seen in 1QJD. The nature of the ligand in the dicarboxylate site was further analyzed in a 1.74 Å resolution structure of avian SQR (PDB code 2H88), confirming the high structural similarity of the ligand and binding site residues in the SQR and flavocytochrome c3 structures (11, 12, 14).Despite the structural information described above, there are still unresolved issues regarding the structure and function of SQRs and QFRs. These include the location of conserved waters, which may form a channel involved in protonation of quinone, and the ability of the Q-site to accommodate different quinones and inhibitors. To further address these issues, we pursued structure-function studies of E. coli SQR. We developed alternative crystallization conditions that provided crystals more reproducibly and diffracting to higher resolution. By exchanging the enzyme into decyl-β-d-maltoside (DM) during purification, it was possible to crystallize the enzyme in the orthorhombic P212121 space group. These crystals routinely diffracted in the 3–3.5 Å resolution range. Co-crystallization with the biochemically well characterized Q-site inhibitor carboxin improved diffraction to 2.1–2.8 Å. This structure shows new features related to the dicarboxylate-binding site of E. coli SQR including a rare cis-peptide bond in SdhA, as found in avian SQR (14), which helps shape the geometry of the active site. Comparisons of the structure with those of SQR binding PCP and SQR with an empty Q-site show how subtle rearrangements of the Q-site accommodate the different inhibitors. The orientation of carboxin in the Q-site differs with computational predictions (16) and with that seen in avian SQR (2FBW). The position of conserved water molecules around the Q-site suggests a new water-mediated proton uptake pathway consistent with recent mutational and biophysical studies (19).  相似文献   
84.
Several studies have shown an immunomodulatory effect of orally administered bovine lactoferrin (LF) in fish, but the process of digestion was not characterized.In the present study, we investigated the fate of bovine LF after oral and anal administration, and studied the appearance of intact LF in the bloodstream and its proteolytic attack during the gastric transit in rainbow trout (Oncorhynchus mykiss) held at 9 °C and 18 °C.Data obtained showed the presence of intact bovine LF in the bloodstream only after anal administration in fish held at 18 °C and the presence of several peptides derived from bovine LF in the gastric content. Immunoblotting analysis showed that only a part of bovine LF-derived peptides reacted with the applied anti-bovine LF antibody. The concentration of intact bovine LF, after 30 min of administration, in the gastric content of fish reared at 18 °C, being extremely low, if any, led us to suspect that the immunoregulatory effect of dietary bovine LF shown in fish by several authors is not due to the intact form but to bioactive fragments, originated by the proteolytic attack during the gastric transit, as demonstrated in higher vertebrates.  相似文献   
85.
Hippocampus dentate gyrus (DG) is characterized by neuronal plasticity processes in adulthood, and polysialylation of NCAM promotes neuronal plasticity. In previous investigations we found that α‐tocopherol increased the PSA‐NCAM‐positive granule cell number in adult rat DG, suggesting that α‐tocopherol may enhance neuronal plasticity. To verify this hypothesis, in the present study, structural remodeling in adult rat DG was investigated under α‐tocopherol supplementation conditions. PSA‐NCAM expression was evaluated by Western blotting, evaluation of PSA‐NCAM‐positive granule cell density, and morphometric analysis of PSA‐NCAM‐positive processes. In addition, the optical density of synaptophysin immunoreactivity and the synaptic profile density, examined by electron microscopy, were evaluated. Moreover, considering that PSA‐NCAM expression has been found to be related to PKCδ activity and α‐tocopherol has been shown to inhibit PKC activity in vitro, Western blotting and immunohistochemistry followed by densitometry were used to analyze PKC. Our results demonstrated that an increase in PSA‐NCAM expression and optical density of DG molecular layer synaptophysin immunoreactivity occurred in α‐tocopherol‐treated rats. Electron microscopy analysis showed that the increase in synaptophysin expression was related to an increase in synaptic profile density. In addition, Western blotting revealed a decrease in phospho‐PKC Pan and phospho‐PKCδ, demonstrating that α‐tocopherol is also able to inhibit PKC activity in vivo. Likewise, immunoreactivity for the active form of PKCδ was lower in α‐tocopherol‐treated rats than in controls, while no changes were found in PKCδ expression. These results demonstrate that α‐tocopherol is an exogenous factor affecting neuronal plasticity in adult rat DG, possibly through PKCδ inhibition. © 2006 Wiley Periodicals, Inc. J Neurobiol, 2006  相似文献   
86.

Background  

Hyaluronic acid (HA) is present in many tissues; its presence in serum may be related to certain inflammatory conditions, tissue damage, sepsis, liver malfunction and some malignancies. In the present work, our goal was to investigate the significance of hyaluronic acid effect on erythrocyte flow properties. Therefore we performed in vitro experiments incubating red blood cells (RBCs) with several HA concentrations. Afterwards, in order to corroborate the pathophysiological significance of the results obtained, we replicated the in vitro experiment with ex vivo RBCs from diagnosed rheumatoid arthritis (RA) patients, a serum HA-increasing pathology.  相似文献   
87.
88.
A simple in situ model of alamethicin-permeabilized isolated rat liver mitochondria was used to investigate the channeling of NADH between mitochondrial malate dehydrogenase (MDH) and NADH:ubiquinone oxidoreductase (complex I). Alamethicin-induced pores in the mitochondrial inner membrane allow effective transport of low molecular mass components such as NAD+/NADH but not soluble proteins. Permeabilized mitochondria demonstrate high rates of respiration in the presence of malate/glutamate and NAD+ due to coupled reaction between MDH and complex I. In the presence of pyruvate and lactate dehydrogenase, an extramitochondrial competitive NADH utilizing system, respiration of permeabilized mitochondria with malate/glutamate and NAD+ was completely abolished. These data are in agreement with the free diffusion of NADH and do not support the suggestion of direct channeling of NADH from MDH to complex I.  相似文献   
89.
Mammalian complex I (NADH:ubiquinone oxidoreductase) exists as a mixture of interconvertible active (A) and de-activated (D) forms. The A-form is capable of NADH:quinone-reductase catalysis, but not the D-form. Complex I from the bacterium Paracoccus denitrificans, by contrast, exists only in the A-form. This bacterial complex contains 32 fewer subunits than the mammalian complex. The question arises therefore if the structural complexity of complex I from higher organisms correlates with its ability to undergo the A/D transition. In the present study, it was found that complex I from the bacterium Escherichia coli and from non-vertebrate organisms (earthworm, lobster, and cricket) did not show the A/D transitions. Vertebrate organisms (carp, frog, chicken), however, underwent similar A/D transitions to those of the well-characterized bovine complex I. Further studies showed that complex I from the lower eukaryotes, Neurospora crassa and Yarrowia lipolytica, exhibited very distinct A/D transitions with much lower activation barriers compared to the bovine enzyme. The A/D transitions of complex I as they relate to structure and regulation of enzymatic activity are discussed.  相似文献   
90.
AIMS: To assess the activity of biostatic agents on the microbial colonization of panel filters. METHODS AND RESULTS: Microfibre glass acrylic filters, both used and unused, were examined for the presence of microorganisms. Test strains were used to verify microbial colonization of filter media. Antimicrobial agents were applied to the filter media and tested for their ability to reduce microbial colonization. The integrity of the panel filters and the antimicrobial activity trends of the filter media treated with antimicrobials were verified. A filtration efficiency test was carried out on the treated filters to evaluate filtration performance. Filters treated with antimicrobials demonstrated markedly less microbial colonization (density and varieties of species), higher filtration efficiency and delayed deterioration of the filter. CONCLUSIONS: The most important results of this study are the demonstration of preservation of the integrity of the filters and the lower release of microorganisms from treated filters. SIGNIFICANCE AND IMPACT OF THE STUDY: These results contribute to the resolution of problems concerning the microbial contamination of panel filters in the heating, ventilating and air-conditioning systems commonly used in the electronic industry, pharmaceutical industry, hospitals and other environments where the absence of contaminating particles and microorganisms is required.  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