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   2449篇
  免费   170篇
  2619篇
  2024年   3篇
  2023年   21篇
  2022年   39篇
  2021年   85篇
  2020年   50篇
  2019年   57篇
  2018年   77篇
  2017年   48篇
  2016年   104篇
  2015年   157篇
  2014年   207篇
  2013年   211篇
  2012年   243篇
  2011年   245篇
  2010年   149篇
  2009年   122篇
  2008年   158篇
  2007年   145篇
  2006年   108篇
  2005年   106篇
  2004年   82篇
  2003年   75篇
  2002年   66篇
  2001年   9篇
  2000年   5篇
  1999年   12篇
  1998年   13篇
  1997年   6篇
  1996年   2篇
  1995年   2篇
  1994年   4篇
  1993年   3篇
  1992年   1篇
  1991年   1篇
  1990年   1篇
  1988年   1篇
  1984年   1篇
排序方式: 共有2619条查询结果,搜索用时 15 毫秒
991.
Clathrin depletion by ribonucleic acid interference (RNAi) impairs mitotic spindle stability and cytokinesis. Depletion of several clathrin-associated proteins affects centrosome integrity, suggesting a further cell cycle function for clathrin. In this paper, we report that RNAi depletion of CHC17 (clathrin heavy chain 17) clathrin, but not the CHC22 clathrin isoform, induced centrosome amplification and multipolar spindles. To stage clathrin function within the cell cycle, a cell line expressing SNAP-tagged clathrin light chains was generated. Acute clathrin inactivation by chemical dimerization of the SNAP-tag during S phase caused reduction of both clathrin and ch-TOG (colonic, hepatic tumor overexpressed gene) at metaphase centrosomes, which became fragmented. This was phenocopied by treatment with Aurora A kinase inhibitor, suggesting a centrosomal role for the Aurora A-dependent complex of clathrin, ch-TOG, and TACC3 (transforming acidic coiled-coil protein 3). Clathrin inactivation in S phase also reduced total cellular levels of ch-TOG by metaphase. Live-cell imaging showed dynamic clathrin recruitment during centrosome maturation. Therefore, we propose that clathrin promotes centrosome maturation by stabilizing the microtubule-binding protein ch-TOG, defining a novel role for the clathrin-ch-TOG-TACC3 complex.  相似文献   
992.
DNA polymerase zeta catalytic subunit REV3 is known to play an important role in the repair of DNA damage induced by cross-linking and methylating agents. Here, we demonstrate that in Arabidopsis (Arabidopsis thaliana), the basic polymerase activity of REV3 is essential for resistance protection against these different types of damaging agents. Interestingly, its processivity is mainly required for resistance to interstrand and intrastrand cross-linking agents, but not alkylating agents. To better define the role of REV3 in relation to other key factors involved in DNA repair, we perform epistasis analysis and show that REV3-mediated resistance to DNA-damaging agents is independent of the replication damage checkpoint kinase ataxia telangiectasia-mutated and rad3-related homolog. REV3 cooperates with the endonuclease MMS and UV-sensitive protein81 in response to interstrand cross links and alkylated bases, whereas it acts independently of the ATP-dependent DNA helicase RECQ4A. Taken together, our data show that four DNA intrastrand cross-link subpathways exist in Arabidopsis, defined by ATP-dependent DNA Helicase RECQ4A, MMS and UV-sensitive protein81, REV3, and the ATPase Radiation Sensitive Protein 5A.The DNA of all living organisms is constantly exposed to damaging factors, and therefore a number of DNA damage repair and bypass mechanisms have evolved. DNA lesions that interfere with the replication machinery constitute a particular challenge for cells (Schröpfer et al., 2014a); if not repaired in a timely manner, such damage can result in the stalling or collapse of replication forks, which in turn can lead to cell death. Furthermore, one-sided double-strand breaks (DSBs) can occur when the replication fork encounters a single-strand break. Lesions within one DNA strand, such as alkylations or DNA intrastrand cross links, can be bypassed by postreplicative repair (PRR), a process that is best understood in yeast (Saccharomyces cerevisiae). This mechanism does not lead to repair of the lesion but prevents fatal long-lasting stalling of the replication fork. PRR can be divided into two branches: the error-prone pathway and the error-free pathway (for review, see Goodman and Woodgate, 2013; Haynes et al., 2015; Jansen et al., 2015). It is known from yeast that both branches of PRR are controlled by the Radiation sensitivity protein6 (Rad6) and Mms-Ubc13 E3 ubiquitin-conjugating enzyme complexes, which ubiquitinate the replicative processivity factor Proliferating Cellular Nuclear Antigen1. The monoubiquitination of PCNA at Lys-164 by Rad6-Rad18 initiates the error-prone pathway, whereas polyubiquitination additionally requires Mms2, Ubc13, and Rad5, and triggers the error-free PRR branch (Hoege et al., 2002; Moldovan et al., 2007; Lee and Myung, 2008). There are two possible competing models postulated for the error-free bypass of lesions at the replication fork, both of which depend on template-switch mechanisms; if the lesion concerns only one of the two sister strands, the undamaged strand can be used as the template for bypassing the lesion. One of the two models features the so-called overshoot synthesis, whereby the newly synthesized strand on the undamaged parental strand is elongated further than the strand blocked by the lesion. Regression of the replication fork then leads to the formation of a special type of four-way junction called a chicken-foot structure. This regression mechanism is thought to be accomplished by helicases, such as the RecQ helicase Bloom Syndrome Protein (BLM) in humans (Croteau et al., 2014). AtRECQ4A is the respective BLM homolog in Arabidopsis (Arabidopsis thaliana), and this enzyme has the ability to regress replication forks in vitro (Hartung et al., 2007, 2008; Schröpfer et al., 2014b). The second error-free subpathway entails invasion of the newly synthesized strand on the blocked sister chromatid into the complementary newly replicated strand on the other sister chromatid. Such a step forms a displacement loop-like structure in which synthesis over the damaged region can occur. In yeast, both error-free pathways are dependent on the multifunctional protein Rad5, which is known to recruit PRR factors and also exhibits helicase activity itself (Blastyák et al., 2007). We previously identified AtRAD5A as a functional Arabidopsis homolog of Rad5 (Chen et al., 2008). Interestingly, AtRAD5A is required for efficient repair by homologous recombination via the synthesis-dependent strand-annealing mechanism, a pathway that in some steps is related to the invasion model of PRR (Mannuss et al., 2010).The error-prone pathway is based on the function of translesion synthesis (TLS) polymerases, which promote replication through DNA lesions (Prakash et al., 2005). In a mechanism termed polymerase switch, the replicative polymerase is exchanged by such a TLS polymerase at a damaged site. After incorporation of a nucleotide opposite the damaged base by the TLS polymerase, a second polymerase switch exchanges the TLS polymerase for the replicative polymerase so that replication can proceed (Prakash and Prakash, 2002; Lehmann et al., 2007). TLS polymerases possess no 5′-3′-exonuclease activity, and therefore act in a potentially mutagenic manner. Nevertheless, depending on the damage incurred and the TLS polymerase used, damage bypass can be error free (Haracska et al., 2000; McCulloch et al., 2004).Polymerases can be divided into at least six families based on their amino acid sequences and crystal structures: A, B, C, D, X, and Y. All of them share the common structure analogous to a right hand grasping DNA with palm, finger, and thumb domains (Steitz, 1999). The amino acid sequences of the finger and thumb domains of different polymerase families are highly variable, whereas the palm domains share high similarity. The palm domain forms the largest part of the polymerase active site and contains highly conserved Asp residues that have been postulated to be involved in the catalytic activity of the enzyme (Joyce and Steitz, 1995; Steitz, 1999).Most TLS polymerases belong to the Y family of polymerases, a class of specially structured enzymes that catalyze replication over damaged templates (Ohmori et al., 2001; Sale et al., 2012). Although some polymerases of the A, B, or X family can also exhibit TLS activity, this is often not their primary function (Prakash et al., 2005). DNA Polymerase Zeta (POLζ) is a B family polymerase and consists of a DNA Polymerase Zeta subunit REV3-REV7 heterodimer, in which REV3 is the catalytic subunit with its accessory subunit and processivity factor REV7 (Nelson et al., 1996). Recent studies in yeast and human cells have shown that POLζ contains two additional subunits, Pol31 and Pol32 in yeast, orthologs to human POLD2 and POLD3, which are known to be accessory subunits of the replicative polymerase POLδ (Johnson et al., 2012; Lee et al., 2014). REV3 contains three regions that are highly conserved between organisms: an N-terminal region, a REV7 binding domain, and a B family-type polymerase domain. The polymerase domain carries the six common conserved regions, I to VI (IV-II-VI-III-I-V), of which I is the most and VI the least conserved region. The A (II), B (III), and C (I) motifs, located within regions I, II, and III (Wong et al., 1988), form the active site of the enzyme, and each harbors an essential Asp residue that coordinates two catalytic metal ions. Deficiency of REV3 in mice is embryo lethal (Bemark et al., 2000; Esposito et al., 2000), and vertebrate cells depleted in REV3 show hypersensitivity to various DNA-damaging agents, including UV and ionizing irradiation, cisplatin, MMS, and mitomycin C (MMC; Sonoda et al., 2003; Sharma and Canman, 2012). In Arabidopsis, rev3 mutants exhibit no obvious phenotype under standard growth conditions, but are hypersensitive to UV-B and gamma irradiation, MMC, MMS, and cisplatin (Sakamoto et al., 2003).Our previous work demonstrated the existence of several different pathways in Arabidopsis involved in repairing the DNA damage induced by cross-linking and methylating agents. These independent pathways are defined by the ATPase RAD5A, the helicase RECQ4A, and MMS and UV-sensitive protein81 (MUS81; Mannuss et al., 2010). The structure-specific endonuclease MUS81 together with its noncatalytic subunit (Mms4 in yeast, Eme1 in Schizosaccharomyces pombe, and MMS4 or Crossover Junction Endonuclease (EME1) in humans and plants) functions in the rescue of stalled replication forks. The enzyme is able to cleave the stalled fork at the lesion site, which leads to a one-sided DSB that is repaired by homologous recombination (HR) to restore the stalled fork (Hanada et al., 2006). We previously showed that mus81 transfer DNA (T-DNA) insertion lines in Arabidopsis are highly sensitive to treatment with MMS, cisplatin, hydroxyurea, ionizing irradiation, and MMC. We also found that AtMUS81 can form a complex with its heterologous binding partners AtEME1A or AtEME1B that is able to process intricate DNA structures, such as nicked Holliday junctions, which might also form at stalled replication forks (Hartung et al., 2006; Geuting et al., 2009).In the current study, we address whether fully functional polymerase activity is required for the repair of DNA damage induced by alkylating and cross-link-inducing agents. Moreover, we sought to clarify whether REV3 cooperates with other key factors identified in Arabidopsis in the repair of these different types of damage. Indeed, it has already been shown that AtREV3 and AtRAD5A do not cooperate in the repair of such DNA damage, confirming independent pathways of error-prone and error-free PRR in plants (Wang et al., 2011). However, as plants, animals, and yeast differ in their DNA cross-link repair machinery (e.g. Mannuss et al., 2010; Knoll et al., 2012; Dangel et al., 2014; Herrmann et al., 2015), it is of particular importance to define the role of REV3 in relation to MUS81 and RECQ4A.  相似文献   
993.
Present on the plasma membrane of most metazoans, caveolae are specialized microdomains implicated in several endocytic and trafficking mechanisms. Caveolins and the more recently discovered cavins are the major protein components of caveolae. Previous studies reported that caveolar invaginations can be induced de novo on the surface of caveolae-negative mammalian cells upon heterologous expression of caveolin-1. However, it remains undocumented whether other components in the transfected cells participate in caveolae formation. To address this issue, we have exploited the protozoan Toxoplasma as a heterologous expression system to provide insights into the minimal requirements for caveogenesis and caveolar endocytosis. Upon expression of caveolin-1, Toxoplasma accumulates prototypical exocytic caveolae ‘precursors’ in the cytoplasm. Toxoplasma expressing caveolin-1 alone, or in conjunction with cavin-1, neither develops surface-located caveolae nor internalizes caveolar ligands. These data suggest that the formation of functional caveolae at the plasma membrane in Toxoplasma and, by inference in all non-mammalian cells, requires effectors other than caveolin-1 and cavin-1. Interestingly, Toxoplasma co-expressing caveolin-1 and cavin-1 displays an impressive spiraled network of membranes containing the two proteins, in the cytoplasm. This suggests a synergistic activity of caveolin-1 and cavin-1 in the morphogenesis and remodeling of membranes, as illustrated for Toxoplasma.  相似文献   
994.

Purpose

We sought to develop placental growth factor as a predictive pharmacodynamic biomarker for motesanib efficacy as first-line therapy in patients with advanced nonsquamous non–small-cell lung cancer.

Experimental Design

Placental growth factor was evaluated at baseline and study week 4 (after 3 weeks treatment) in an exploratory analysis of data from a randomized phase 2 study of motesanib 125 mg once daily plus carboplatin/paclitaxel and in a prespecified analysis of data from a randomized, double-blind phase 3 study of motesanib 125 mg once daily plus carboplatin/paclitaxel vs placebo plus carboplatin/paclitaxel (MONET1). Associations between fold-change from baseline in placental growth factor and overall survival were evaluated using Cox proportional hazards models.

Results

In the phase 2 study, serum placental growth factor increased from baseline a mean 2.8-fold at study week 4. Patients with ≥2.2-fold change from baseline in placental growth factor (n = 18) had significantly longer overall survival than those with <2.2-fold change (n = 19; 22.9 vs 7.9 months; hazard ratio, 0.30; 95% CI, 0.12–0.74; P = 0.009). Consequently, placental growth factor was investigated as a pharmacodynamic biomarker in the phase 3 MONET1 study. There was no association between log-transformed placental growth factor fold-change from baseline to week 4 (continuous variable) and overall survival (hazard ratio, 0.98; 95% CI, 0.79–1.22; P = 0.868). MONET1 did not meet its primary endpoint of overall survival. Likewise, median overall survival was similar among patients with ≥2.0-fold change in placental growth factor (n = 229) compared with <2.0-fold change (n = 127; 14.8 vs 13.8 months; hazard ratio, 0.88; 95% CI, 0.67–1.15, P = 0.340).

Conclusions

Our results illustrate the challenges of successfully translating phase 2 biomarker results into phase 3 studies.

Trial Registration

ClinicalTrials.gov NCT00460317, NCT00369070  相似文献   
995.
996.
Glycine max is one of the major sources of phytochemicals, in particular of isoflavones, a class of phytoestrogens with ascertained beneficial effects on human health. In the present study, in vitro callus production from soybean hypocotyl seedling explants and cell suspensions were optimized. Time-courses having 20, 40 and 60 g/L of initial cell inoculum were performed to determine the concentration most suitable for isoflavone production. The amount of total polyphenols and total flavonoids as well as the antioxidant capacity of both cell and culture media fractions were measured by means of spectrophotometric methods. The levels of aglycone and glycosylated isoflavones (didzein, genistein, glycitein, didzin, genistin, glycitin), as well as of ferulic acid, vanillic acid and vanillin, were determined by HPLC–DAD. On average, 93.5 % of the total (cells plus media) isoflavones in soybean cell suspensions were detected as aglycones. Concentrated cell cultures as well as industrial soybean seed extracts were enzymatically hydrolyzed to release the aglycones and their metabolic profiles were analysed by HPLC–DAD. In contrast to cell suspensions, in undigested seed extract the aglycon form represented only 16.8 % of the total isoflavones amount. After enzymatic treatment, the antioxidant capacity increased by 30 and 33 %, respectively, in concentrated cell and seed extracts, demonstrating the presence of a larger amount of bioactive metabolites after digestion. At the present extraction conditions, soybean concentrated cell suspensions yielded 5.8-fold more total isoflavones (mostly in the free form) than seed extracts, leading to hypothesise their possible use as ingredients for cosmetic and nutraceutical applications.  相似文献   
997.
Mitophagy, the selective removal of damaged or excess mitochondria by autophagy, is an important process in cellular homeostasis. The outer mitochondrial membrane (OMM) proteins NIX, BNIP3, FUNDC1, and Bcl2‐L13 recruit ATG8 proteins (LC3/GABARAP) to mitochondria during mitophagy. FKBP8 (also known as FKBP38), a unique member of the FK506‐binding protein (FKBP) family, is similarly anchored in the OMM and acts as a multifunctional adaptor with anti‐apoptotic activity. In a yeast two‐hybrid screen, we identified FKBP8 as an ATG8‐interacting protein. Here, we map an N‐terminal LC3‐interacting region (LIR) motif in FKBP8 that binds strongly to LC3A both in vitro and in vivo. FKBP8 efficiently recruits lipidated LC3A to damaged mitochondria in a LIR‐dependent manner. The mitophagy receptors BNIP3 and NIX in contrast are unable to mediate an efficient recruitment of LC3A even after mitochondrial damage. Co‐expression of FKBP8 with LC3A profoundly induces Parkin‐independent mitophagy. Strikingly, even when acting as a mitophagy receptor, FKBP8 avoids degradation by escaping from mitochondria. In summary, this study identifies novel roles for FKBP8 and LC3A, which act together to induce mitophagy.  相似文献   
998.
The aim of this study was to evaluate the safety and probiotic potential characteristics of ten Lactobacillus spp. strains (Lactobacillus fermentum SJRP30, Lactobacillus casei SJRP37, SJRP66, SJRP141, SJRP145, SJRP146, and SJRP169, and Lactobacillus delbrueckii subsp. bulgaricus SJRP50, SJRP76, and SJRP149) that had previously been isolated from water buffalo mozzarella cheese. The safety of the strains was analyzed based on mucin degradation, hemolytic activity, resistance to antibiotics and the presence of genes encoding virulence factors. The in vitro tests concerning probiotic potential included survival under simulated gastrointestinal (GI) tract conditions, intestinal epithelial cell adhesion, the presence of genes encoding adhesion, aggregation and colonization factors, antimicrobial activity, and the production of the β-galactosidase enzyme. Although all strains presented resistance to several antibiotics, the resistance was limited to antibiotics to which the strains had intrinsic resistance. Furthermore, the strains presented a limited spread of genes encoding virulence factors and resistance to antibiotics, and none of the strains presented hemolytic or mucin degradation activity. The L. delbrueckii subsp. bulgaricus strains showed the lowest survival rate after exposure to simulated GI tract conditions, whereas all of the L. casei and L. fermentum strains showed good survivability. None of the tested lactobacilli strains presented bile salt hydrolase (BSH) activity, and only L. casei SJRP145 did not produce the β-galactosidase enzyme. The strains showed varied levels of adhesion to Caco-2 cells. None of the cell-free supernatants inhibited the growth of pathogenic target microorganisms. Overall, L. fermentum SJRP30 and L. casei SJRP145 and SJRP146 were revealed to be safe and to possess similar or superior probiotic characteristics compared to the reference strain L. rhamnosus GG (ATCC 53103).  相似文献   
999.
The aim of this work was to assess whether "modeled microgravity" affects cell response to ionizing radiation, increasing the risk associated with radiation exposure. Lymphoblastoid TK6 cells were irradiated with various doses of gamma rays and incubated for 24 h in a modeled microgravity environment obtained by the Rotating Wall Vessel bioreactor. Cell survival, induction of apoptosis and cell cycle alteration were compared in cells irradiated and then incubated in 1g or modeled microgravity conditions. Modulation of genomic damage induced by ionizing radiation was evaluated on the basis of HPRT mutant frequency and the micronucleus assay. A significant reduction in apoptotic cells was observed in cells incubated in modeled microgravity after gamma irradiation compared with cells maintained in 1g. Moreover, in irradiated cells, fewer G2-phase cells were found in modeled microgravity than in 1g, whereas more G1-phase cells were observed in modeled microgravity than in 1g. Genomic damage induced by ionizing radiation, i.e. frequency of HPRT mutants and micronucleated cells, increased more in cultures incubated in modeled microgravity than in 1g. Our results indicate that modeled microgravity incubation after irradiation affects cell response to ionizing radiation, reducing the level of radiation-induced apoptosis. As a consequence, modeled microgravity increases the frequency of damaged cells that survive after irradiation.  相似文献   
1000.
The causative agent of Legionnaires'' pneumonia, Legionella pneumophila, colonizes diverse environmental niches, including biofilms, plant material, and protozoa. In these habitats, myo-inositol hexakisphosphate (phytate) is prevalent and used as a phosphate storage compound or as a siderophore. L. pneumophila replicates in protozoa and mammalian phagocytes within a unique “Legionella-containing vacuole.” The bacteria govern host cell interactions through the Icm/Dot type IV secretion system (T4SS) and ∼300 different “effector” proteins. Here we characterize a hitherto unrecognized Icm/Dot substrate, LppA, as a phytate phosphatase (phytase). Phytase activity of recombinant LppA required catalytically essential cysteine (Cys231) and arginine (Arg237) residues. The structure of LppA at 1.4 Å resolution revealed a mainly α-helical globular protein stabilized by four antiparallel β-sheets that binds two phosphate moieties. The phosphates localize to a P-loop active site characteristic of dual specificity phosphatases or to a non-catalytic site, respectively. Phytate reversibly abolished growth of L. pneumophila in broth, and growth inhibition was relieved by overproduction of LppA or by metal ion titration. L. pneumophila lacking lppA replicated less efficiently in phytate-loaded Acanthamoeba castellanii or Dictyostelium discoideum, and the intracellular growth defect was complemented by the phytase gene. These findings identify the chelator phytate as an intracellular bacteriostatic component of cell-autonomous host immunity and reveal a T4SS-translocated L. pneumophila phytase that counteracts intracellular bacterial growth restriction by phytate. Thus, bacterial phytases might represent therapeutic targets to combat intracellular pathogens.  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