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   1425篇
  免费   82篇
  国内免费   6篇
  1513篇
  2023年   8篇
  2022年   19篇
  2021年   55篇
  2020年   17篇
  2019年   25篇
  2018年   39篇
  2017年   27篇
  2016年   44篇
  2015年   54篇
  2014年   57篇
  2013年   95篇
  2012年   72篇
  2011年   96篇
  2010年   55篇
  2009年   49篇
  2008年   71篇
  2007年   75篇
  2006年   45篇
  2005年   53篇
  2004年   50篇
  2003年   41篇
  2002年   33篇
  2001年   23篇
  2000年   14篇
  1999年   22篇
  1998年   10篇
  1997年   10篇
  1996年   10篇
  1994年   8篇
  1992年   17篇
  1991年   17篇
  1990年   16篇
  1989年   14篇
  1988年   14篇
  1987年   17篇
  1986年   12篇
  1985年   18篇
  1984年   14篇
  1983年   15篇
  1982年   10篇
  1981年   9篇
  1980年   9篇
  1979年   16篇
  1978年   12篇
  1977年   9篇
  1976年   11篇
  1974年   15篇
  1973年   7篇
  1971年   9篇
  1969年   10篇
排序方式: 共有1513条查询结果,搜索用时 15 毫秒
31.
O2 sensing in diverse protozoa depends on the prolyl 4 hydroxylation of Skp1 and modification of the resulting hydroxyproline with a series of five sugars. In yeast, plants, and animals, Skp1 is associated with F-box proteins. The Skp1–F-box protein heterodimer can, for many F-box proteins, dock onto cullin-1 en route to assembly of the Skp1–cullin-1–F-box protein–Rbx1 subcomplex of E3SCFUb ligases. E3SCFUb ligases conjugate Lys48-polyubiquitin chains onto targets bound to the substrate receptor domains of F-box proteins, preparing them for recognition by the 26S proteasome. In the social amoeba Dictyostelium, we found that O2 availability was rate-limiting for the hydroxylation of newly synthesized Skp1. To investigate the effect of reduced hydroxylation, we analyzed knockout mutants of the Skp1 prolyl hydroxylase and each of the Skp1 glycosyltransferases. Proteomic analysis of co-immunoprecipitates showed that wild-type cells able to fully glycosylate Skp1 had a greater abundance of an SCF complex containing the cullin-1 homolog CulE and FbxD, a newly described WD40-type F-box protein, than the complexes that predominate in cells defective in Skp1 hydroxylation or glycosylation. Similarly, the previously described FbxA–Skp1CulA complex was also more abundant in glycosylation-competent cells. The CulE interactome also included higher levels of proteasomal regulatory particles when Skp1 was glycosylated, suggesting increased activity consistent with greater association with F-box proteins. Finally, the interactome of FLAG-FbxD was modified when it harbored an F-box mutation that compromised Skp1 binding, consistent with an effect on the abundance of potential substrate proteins. We propose that O2-dependent posttranslational glycosylation of Skp1 promotes association with F-box proteins and their engagement in functional E3SCFUb ligases that regulate O2-dependent developmental progression.Timely protein degradation is a cornerstone of cell cycling and the regulation of numerous physiological and developmental processes. Eukaryotes have evolved an extensive array of polyubiquitination enzymes to tag proteins on a protein-by-protein basis as a recognition marker for degradation in the 26S proteasome. The cullin-RING ubiquitin ligases (CRLs)1 are a prominent subgroup of these enzymes (1) and consist of an E3 architecture that includes a substrate receptor, an adaptor (in most cases), the cullin scaffold, the RING protein, and an exchangeable E2 ubiquitin donor that has been charged with ubiquitin (Ub) by an E1 enzyme. The first discovered and still prototypic example is the CRL1 class (2), also referred to as SCF on account of the names of its founding subunits, Skp1, cullin-1, and F-box proteins (FBPs). The CRL1 (or SCF) complexes utilize FBPs as substrate receptors, Skp1 as the adaptor linking the FBP to the N-terminal region of cullin-1 (Cul1), and Rbx1 as the RING protein that tethers the E2 Ub donor to the Cul1 C-terminal region (see Fig. 2B). CRL1s can be activated by neddylation of Cul1 by a Nedd8-specific E2, which mobilizes Rbx1 to afford rotational flexibility of the E2 and displaces the inhibitor Cand1, permitting docking of the Skp1–FBP heterodimer (35). Deneddylation mediated by the eight-subunit COP9 signalosome is required for in vivo activity, suggesting that Cand1 serves as a substrate exchange factor to allow for re-equilibration of SCF complexes from preexisting subunits. Each reaction cycle requires the exchange of a new E2-Ub and typically assembles a K48-linked polyUb chain that is recognized by the proteasome. Substrate specificity is conferred by FBPs, a gene family that numbers 69 in humans, 20 in budding yeast, 300 in Caenorhabditis elegans, and ∼800 in Arabidopsis. Some characterized FBPs can recognize perhaps a dozen or more substrates, and the coding of recognition and the meaning of their control by the same FBP is under intense investigation (6). Recognition is often activated by posttranslational modification of the substrate (often phosphorylation). Regulation of SCF Ub ligases has centered on the neddylation cycle, which potentially influences all seven known CRLs. Regulation of Skp1, investigated in this paper, would be specific to CRLs possessing Skp1, which include CRL1 and possibly the minor class CRL7 (7).Open in a separate windowFig. 2.Skp1 modification pathway and global analysis of Skp1 interactions. A, Skp1 is sequentially modified by the indicated enzymes (in blue), resulting in the formation of a pentasaccharide at Pro143. B, model of the SCF complex in the context of the overall E3 Ub ligase, from studies in yeast, plants, and animals. Catalysis involves transfer of Ub from an exchangeable Ub-E2 conjugate to the substrate. Removal of Nedd8 by the COP9 signalosome facilitates binding of Cand1 to Cul1, which inhibits binding of Skp1 to Cul1. C, D, vegetative (growth stage) cells were filter-lysed, and a cytosolic fraction prepared via ultracentrifugation was chromatographed on a Superose 12 gel filtration column. Fractions were analyzed via Western blotting (representative examples are shown in C) followed by densitometry (D). The elution position of free Skp1 from a separate trial is indicated.The basic SCF model is thought to be widespread among eukaryotes but has been extensively studied only in fungi/yeasts, plants, and animals. The broad phylogeny represented by protists includes many benign and pathogenic unicellular organisms of great economic, health, and environmental impact. Emerging evidence reveals that Skp1 in some of these groups is subject to a novel form of prolyl 4(trans)-hydroxylation and complex glycosylation (8). The roles of these Skp1 modifications have been most studied in the social amoeba Dictyostelium, which undergoes a starvation-induced developmental program during which individual amoebae chemotactically aggregate into an initial mound that then elongates into a migratory slug. Under appropriate conditions, the slug reorganizes to form a fruiting body consisting of a ball of spores supported by a vertical cellular stalk. The slug-to-fruit switch, referred to as culmination, and sporulation are regulated by checkpoints that are sensitive to multiple factors, including O2 (911). Functional studies of Dictyostelium Skp1 hydroxylation and glycosylation reveal roles in regulating the O2 dependence of culmination and sporulation (1214). For example, wild-type (wt) cells require 7% to 10% O2 and phyA requires 18% to 21% O2 in order to achieve 50% spore formation (a quantitative measure of fruiting body formation), whereas glycosylation mutants exhibit a complex pattern of intermediate requirements (13). In addition, at 21% O2, phyA cells require an additional 3 to 4 h to complete development relative to their wt counterparts (14). In the apicomplexan Toxoplasma gondii, PhyA is also required for Skp1 glycosylation, and phyA parasites are deficient in proliferation, especially at low O2 (15).The idea that O2 availability is rate limiting for Skp1 modification was originally based on the observation that the Dictyostelium phyA phenotype mimics that of wt cells in low O2 (9). However, the majority of Skp1 is hydroxylated and glycosylated in wt cells even at low O2 levels where culmination is blocked or delayed. Further analysis of a submerged development model, in which terminal development depended on an atmosphere of 70% to 100% O2 in order to overcome the diffusion barrier posed by the water layer, showed that at atmospheric O2 levels of 5% to 21% where sporulation was blocked, unmodified Skp1 accumulated to a higher level than at permissive O2 levels (10). As Skp1 modifications are thought to be irreversible, this likely resulted from slow hydroxylation of newly synthesized Skp1. To address this in a more physiological setting, we investigated nascent Skp1 directly using metabolic labeling with [35S]Met/Cys and verified that the rate of hydroxylation of newly synthesized Skp1 polypeptide was indeed inversely proportional to O2 levels, which makes PhyA-mediated hydroxylation of Skp1 an excellent candidate for the primary O2 sensor for culmination.These modifications of Skp1 are of interest as a novel mechanism regulating the SCF ligase. Previously, we showed that hydroxylation and glycosylation of Dictyostelium Skp1 affect its conformation and promote binding to a soluble FBP, guinea pig Fbs1, in studies of purified proteins (16). Here we show that Dictyostelium Skp1 is indeed a subunit of a canonical SCF complex, as expected. The significance of undermodified Skp1 was examined via interactome analysis of Skp1 isoforms that accumulate in modification pathway mutants. Our findings revealed a lower abundance of SCF complexes than in wt cells, suggesting that Skp1 modification may promote SCF assembly and E3SCFUb ligase activities that control timely turnover of select proteins involved in developmental progression.  相似文献   
32.
Acclimated activated sludge was examined for its ability to degrade malathion with and without the presence of glucose as a potential cometabolite substrate. In this study, a packed-bed reactor (PBR) using three kinds of biofilm carriers was employed for efficient degradation of malathion. The results obtained indicate that microorganisms tested were able to degrade malathion. The observed degradation rate of the pesticide in the presence of glucose was the same as without glucose. The activated sludge was found to be able to use malathion as the sole phosphorus source. In contrast, the degradation ability of the activated sludge was lost when the pesticide was used as the sole source of sulfur. The degradation capacity of the PBR was higher than the performance obtained with the batch reactor. The reactor packed with crushed olive kernels exhibited the best performance, allowing a total removal of malathion (10 mg/dm3) within 12 h.  相似文献   
33.
34.
A hitherto unknown hydroxy acid has been isolated from Blepharis sindica seed oil has been characterized as 9-hydroxydodecanoic acid by IR, NMR and mass spectral studies. The structure of this acid was further supported by its chemical transformations.  相似文献   
35.
Cyclodextrin glucanotransferase (CGTase) activity was observed when the bacterium was grown in the medium at various initial pH values, containing carbon, nitrogen, phosphorus and mineral salt sources at 50 °C for 24 h in the shake flasks. The optimisation of this growth medium was carried out using response surface methodology. The design contains a total of 32 experimental trials involving 10 star points and 6 replicates at the centre points. The design was employed by selecting sago starch, peptone from casein, K2HPO4, CaCl2 and initial pH as five independent variables in this study. The optimal calculated values of tested variables for maximal production of CGTase were found to be comprised of: sago starch, 16.02 g/l; peptone from casein, 20 g/l; K2HPO4, 1.4 g/l; CaCl2, 0.2 g/l and initial pH, 7.54 with a predicted CGTase activity of 14.20 U/ml. These predicted optimal parameters were tested in the laboratory and the final CGTase activity obtained was very close to the predicted value at 14.80 U/ml.  相似文献   
36.
In insects, developmental responses are organ- and tissue-specific. In previous studies of insect midgut cells in primary tissue cultures, growth-promoting and differentiation factors were identified from the growth media, hemolymph, and fat body. Recently, it was determined that the mitogenic effect of a Manduca sexta fat body extract on midgut stem cells of Heliothis virescens was due to the presence of monomeric alpha-arylphorin. Here we report that in primary midgut cell cultures, this same arylphorin stimulates stem cell proliferation in the lepidopterans M. sexta and Spodoptera littoralis, and in the beetle Leptinotarsa decemlineata. Studies using S. littoralis cells confirm that the mitogenic effect is due to free alpha-arylphorin subunits. In addition, feeding artificial diets containing arylphorin increased the growth rates of several insect species. When tested against continuous cell lines, including some with midgut and fat body origins, arylphorin had no effect; however, a cell line derived from Lymantria dispar fat body grew more rapidly in medium containing a chymotryptic digest of arylphorin.  相似文献   
37.
The production of extracellular pullulanase by Bacillus licheniformis NRC22 was investigated using different fermentation modes. In batch culture maximal enzyme activity of 18 U/ml was obtained after 24 h of growth. In continuous fermentation by the free cells, maximal reactor productivity (4.15 KU/l/h) with enzyme concentration of 14.8 U/ml and specific productivity of 334.9 U/g wet cells/h was attained at a dilution rate of 0.28/h, over a period of 25 days. B. licheniformis NRC22 cells were immobilized on Ca-alginate. The immobilization conditions with respect to matrix concentration and cell load was optimized for maximal enzyme production. In repeated batch operation, the activity of the immobilized cells was stable during the 10 cycles and the activity remained between 9.8 and 7.7 U/ml. Continuous production of pullulanase by the immobilized cells was investigated in a packed–bed reactor. Maximal reactor productivity (7.0 KU/h) with enzyme concentration of 16.8 U/ml and specific productivity of 131.64 U/g wet cells/h was attained at dilution rate of 0.42/h. The enzyme activity in the effluent started to decline gradually to the level of 8.7 U/ml after 25 days of the operation.  相似文献   
38.
The mTOR kinase inhibitor rapamycin (sirolimus) is a drug with potent immunosuppressive and antiproliferative properties. We found that rapamycin induces the TGFβ/Smad signaling cascade in rat mesangial cells (MC) as depicted by the nuclear translocation of phospho-Smads 2, -3 and Smad-4, respectively. Concomitantly, rapamycin increases the nuclear DNA binding of receptor (R)- and co-Smad proteins to a cognate Smad-binding element (SBE) which in turn causes an increase in profibrotic gene expression as exemplified by the connective tissue growth factor (CTGF) and plasminogen activator inhibitor 1 (PAI-1). Using small interfering (si)RNA we demonstrate that Smad 2/3 activation by rapamycin depends on its endogenous receptor FK binding protein 12 (FKBP12). Mechanistically, Smad induction by rapamycin is initiated by an increase in active TGFβ1 as shown by ELISA and by the inhibitory effects of a neutralizing TGFβ antibody. Using an activin receptor-like kinase (ALK)-5 inhibitor and by siRNA against the TGFβ type II receptor (TGFβ-RII) we furthermore demonstrate a functional involvement of both types of TGFβ receptors. However, rapamycin did not compete with TGFβ for TGFβ-receptor binding as found in radioligand-binding assay. Besides SB203580, a specific inhibitor of the p38 MAPK, the reactive oxygen species (ROS) scavenger N-acetyl-cysteine (NAC) and a cell-permeable superoxide dismutase (SOD) mimetic strongly abrogated the stimulatory effects of rapamycin on Smad 2 and 3 phosphorylation. Furthermore, the rapid increase in dichlorofluorescein (DCF) formation implies that rapamycin mainly acts through ROS. In conclusion, activation of the profibrotic TGFβ/Smad signaling cascade accompanies the immunosuppressive and antiproliferative actions of rapamycin.  相似文献   
39.
40.
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