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   1030篇
  免费   79篇
  1109篇
  2023年   2篇
  2022年   9篇
  2021年   20篇
  2020年   13篇
  2019年   11篇
  2018年   23篇
  2017年   3篇
  2016年   13篇
  2015年   35篇
  2014年   52篇
  2013年   85篇
  2012年   92篇
  2011年   67篇
  2010年   49篇
  2009年   59篇
  2008年   90篇
  2007年   93篇
  2006年   72篇
  2005年   80篇
  2004年   57篇
  2003年   69篇
  2002年   57篇
  2001年   9篇
  2000年   4篇
  1999年   7篇
  1998年   8篇
  1997年   6篇
  1996年   4篇
  1995年   2篇
  1994年   3篇
  1993年   3篇
  1992年   1篇
  1991年   1篇
  1990年   1篇
  1987年   1篇
  1982年   2篇
  1980年   1篇
  1979年   1篇
  1978年   2篇
  1977年   1篇
  1931年   1篇
排序方式: 共有1109条查询结果,搜索用时 15 毫秒
101.
Drug resistance in Mycobacterium tuberculosis is a global problem, with major consequences for treatment and public health systems. As the emergence and spread of drug‐resistant tuberculosis epidemics is largely influenced by the impact of the resistance mechanism on bacterial fitness, we wished to investigate whether compensatory evolution occurs in drug‐resistant clinical isolates of M. tuberculosis. By combining information from molecular epidemiology studies of drug‐resistant clinical M. tuberculosis isolates with genetic reconstructions and measurements of aminoglycoside susceptibility and fitness in Mycobacterium smegmatis, we have reconstructed a plausible pathway for how aminoglycoside resistance develops in clinical isolates of M. tuberculosis. Thus, we show by reconstruction experiments that base changes in the highly conserved A‐site of 16S rRNA that: (i) cause aminoglycoside resistance, (ii) confer a high fitness cost and (iii) destabilize a stem‐loop structure, are associated with a particular compensatory point mutation that restores rRNA secondary structure and bacterial fitness, while maintaining to a large extent the drug‐resistant phenotype. The same types of resistance and associated mutations can be found in M. tuberculosis in clinical isolates, suggesting that compensatory evolution contributes to the spread of drug‐resistant tuberculosis disease.  相似文献   
102.
Singh ND  Arndt PF  Petrov DA 《Genetics》2005,169(2):709-722
Mutation is the underlying force that provides the variation upon which evolutionary forces can act. It is important to understand how mutation rates vary within genomes and how the probabilities of fixation of new mutations vary as well. If substitutional processes across the genome are heterogeneous, then examining patterns of coding sequence evolution without taking these underlying variations into account may be misleading. Here we present the first rigorous test of substitution rate heterogeneity in the Drosophila melanogaster genome using almost 1500 nonfunctional fragments of the transposable element DNAREP1_DM. Not only do our analyses suggest that substitutional patterns in heterochromatic and euchromatic sequences are different, but also they provide support in favor of a recombination-associated substitutional bias toward G and C in this species. The magnitude of this bias is entirely sufficient to explain recombination-associated patterns of codon usage on the autosomes of the D. melanogaster genome. We also document a bias toward lower GC content in the pattern of small insertions and deletions (indels). In addition, the GC content of noncoding DNA in Drosophila is higher than would be predicted on the basis of the pattern of nucleotide substitutions and small indels. However, we argue that the fast turnover of noncoding sequences in Drosophila makes it difficult to assess the importance of the GC biases in nucleotide substitutions and small indels in shaping the base composition of noncoding sequences.  相似文献   
103.
The C terminus of the herpes simplex virus type 1 origin-binding protein, UL9ct, interacts directly with the viral single-stranded DNA-binding protein ICP8. We show that a 60-amino acid C-terminal deletion mutant of ICP8 (ICP8ΔC) also binds very strongly to UL9ct. Using small angle x-ray scattering, the low resolution solution structures of UL9ct alone, in complex with ICP8ΔC, and in complex with a 15-mer double-stranded DNA containing Box I of the origin of replication are described. Size exclusion chromatography, analytical ultracentrifugation, and electrophoretic mobility shift assays, backed up by isothermal titration calorimetry measurements, are used to show that the stoichiometry of the UL9ct-dsDNA15-mer complex is 2:1 at micromolar protein concentrations. The reaction occurs in two steps with initial binding of UL9ct to DNA (Kd ∼ 6 nm) followed by a second binding event (Kd ∼ 0.8 nm). It is also shown that the stoichiometry of the ternary UL9ct-ICP8ΔC-dsDNA15-mer complex is 2:1:1, at the concentrations used in the different assays. Electron microscopy indicates that the complex assembled on the extended origin, oriS, rather than Box I alone, is much larger. The results are consistent with a simple model whereby a conformational switch of the UL9 DNA-binding domain upon binding to Box I allows the recruitment of a UL9-ICP8 complex by interaction between the UL9 DNA-binding domains.The initiation of DNA replication for most double-stranded DNA (dsDNA)6 viral genomes begins with the recognition of the origin by specific origin-binding proteins. The herpes simplex virus type 1 (HSV-1) genome encodes seven proteins required for origin-dependent DNA replication. These are the DNA polymerase (UL30) and its accessory protein (UL42), a heterotrimeric helicase-primase complex (UL5, UL8, and UL52), the single-stranded DNA-binding protein (ICP8 or UL29), and the origin-binding protein (UL9) (reviewed in Ref. 1). HSV-1 contains three functional origins, oriL and two copies of oriS. OriS, which is about 80 bp in length, consists of three UL9 recognition sites, in Boxes I, II, and III, which are arranged in two overlapping palindromes (2). Box I and Box III are part of an evolutionarily conserved palindrome that forms a stable hairpin in single-stranded DNA, which may be important in the origin rearrangement (3) during initiation of replication. Box I and II are separated by an AT-rich spacer sequence, which varies in length and nucleotide composition between the different members of the α-herpesvirus subfamily (2, 46).UL9 is a homodimer in solution, and EM studies, with UL9 bound to oriS, indicate the existence of a dimer or pair of dimers assembled on oriS (7). Several reports indicate that UL9 can physically interact not only with ICP8 (8) but also with other members of the HSV-1 replication complex, including UL8 (9) and UL42 (10). Thus UL9 functions as a docking protein to recruit these essential replication proteins to the viral origins. ICP8 stimulates the helicase activity of UL9 (11, 12) and binds to its C-terminal 27-aa residues (13). In the presence of ICP8, UL9 will open dsDNA containing Box I, leading to a conformational change in the origin, thus facilitating unwinding (1416). As stated above, the changes in DNA conformation in the complete oriS may be more complex (3). Recently, it has been suggested that single-stranded oriS folds into a unique and evolutionarily conserved conformation, oriS*, which is stably bound by UL9. oriS* contains a hairpin formed by complementary base pairing between Box I and Box III in oriS (17). UL9, in the presence of the single-stranded DNA-binding protein ICP8, can convert an 80-bp double-stranded minimal oriS fragment to oriS* and form a UL9-oriS* complex. The formation of a UL9-oriS* complex requires ATP hydrolysis (18). Therefore, the UL9-oriS* complex may serve as an assembly site for the herpesvirus replisome. Macao et al. (3) proposed a model in which full-length UL9 would be required to adopt a different conformation when binding to oriS or oriS*. The implication is that UL9 partially unwinds and introduces a hairpin into the origin of replication and that the formation of oriS* is aided, in some way, by ICP8 and requires ATP hydrolysis. Macao et al. (3) suggest that the length of the single-stranded tail of the probe DNA determines the stoichiometry of the UL9-DNA complex. oriS may bind two molecules of UL9, whereas oriS* may only bind one because the hairpin formation prevents the second interaction.Photo-cross-linking studies have shown that, although the UL9 protein binds Box I as a dimer, only one of the two monomers contacts Box I, suggesting that the C terminus of UL9 undergoes a conformational change upon binding to Box I (19). The results reported here are consistent with this observation. To date there is no three-dimensional structural information available on the full-length UL9 or either of the functionally characterized (helicase and DNA binding) domains. The ability to adopt different conformations and a tendency to proteolytic degradation may be responsible for this. It has been shown that UL9 binds with very high specificity to the Box I through its DNA-binding domain, consisting of the C-terminal 317 aa (UL9ct) (20, 21). Although the importance of the binding between UL9ct and oriS for the viral life cycle is well established, the mechanism behind this interaction still remains unclear. Even though UL9ct exists as a monomer in solution, uncertainty remains as to whether one or two molecules bind to a single Box I recognition sequence. Some reports have suggested that one UL9ct molecule binds to a single copy of the sequence (2224), whereas others have proposed that UL9ct forms a dimer when bound to DNA (25, 26). This apparent difference may well result from the different protein concentrations used in different assays/experiments, which in turn highlights the difficulty of translating in vitro equilibrium experiments into cellular nonequilibrium situations.A few years ago, the crystal structure of a 60-residue C-terminal deletion mutant of ICP8 (ICP8ΔC) was determined to 3 Å resolution (Protein Data Bank code 1URJ (27)). The structure of ICP8ΔC consists of a large N-terminal domain (aa 9–1038) and a smaller entirely helical C-terminal domain (aa 1049–1120) connected to the N-terminal domain by a disordered linker (aa 1038–1049) spanning around 18 Å in the crystal structure. ICP8 preferentially binds ssDNA over dsDNA in a nonsequence-specific and cooperative manner (28). ICP8 is a zinc metalloprotein containing one zinc atom per molecule, which is coordinated by three cysteines (Cys-499, Cys-502, and Cys-510) and a histidine (His-512) (27).In this study, we show that the 60-amino acid C-terminal deletion of ICP8 (ICP8ΔC) binds strongly to UL9ct. We present three low resolution structures in solution using small angle x-ray scattering as follows: that of the UL9ct alone, in complex with ICP8ΔC, and in complex with a 15-mer dsDNA (dsDNA15-mer) containing the Box I sequence. Using these data and a variety of biophysical techniques, we demonstrate that the stoichiometries of the UL9ct-dsDNA15-mer and UL9ct-ICP8ΔC-dsDNA15-mer complexes are 2:1 and 2:1:1, respectively, at the micromolar protein concentrations used in this study. Using EM we visualize the assembly of the ICP8ΔC-UL9ct complex on oriS and estimate the size of the complex.  相似文献   
104.
105.
Mutual dependence of MDM2 and MDMX in their functional inactivation of p53   总被引:12,自引:0,他引:12  
MDMX, an MDM2-related protein, has emerged as yet another essential negative regulator of p53 tumor suppressor, since loss of MDMX expression results in p53-dependent embryonic lethality in mice. However, it remains unknown why neither homologue can compensate for the loss of the other. In addition, results of biochemical studies have suggested that MDMX inhibits MDM2-mediated p53 degradation, thus contradicting its role as defined in gene knockout experiments. Using cells deficient in either MDM2 or MDMX, we demonstrated that these two p53 inhibitors are in fact functionally dependent on each other. In the absence of MDMX, MDM2 is largely ineffective in down-regulating p53 because of its extremely short half-life. MDMX renders MDM2 protein sufficiently stable to function at its full potential for p53 degradation. On the other hand, MDMX, which is a cytoplasmic protein, depends on MDM2 to redistribute into the nucleus and be able to inactivate p53. We also showed that MDMX, when exceedingly overexpressed, inhibits MDM2-mediated p53 degradation by competing with MDM2 for p53 binding. Our findings therefore provide a molecular basis for the nonoverlapping activities of these two p53 inhibitors previously revealed in genetic studies.  相似文献   
106.
The Poland–Fixman–Freire formalism was adapted for modeling of calorimetric DNA melting profiles, and applied to plasmid pBR 322 and long random sequences. We studied the influence of the difference (HGC?HAT) between the helix‐coil transition enthalpies of AT and GC base pairs on the calorimetric melting profile and on normalized calorimetric melting profile. A strong alteration of DNA calorimetrical profile with HGC?HAT was demonstrated. In contrast, there is a relatively slight change in the normalized profiles and in corresponding ordinary (optical) normalized differential melting curves (DMCs). For fixed HGC?HAT, the average relative deviation (S) between DMC and normalized calorimetric profile, and the difference between their melting temperatures (Tcal?Tm) are weakly dependent on peculiarities of the multipeak fine structure of DMCs. At the same time, both the deviation S and difference (Tcal?Tm) enlarge with the temperature melting range of the helix‐coil transition. It is shown that the local deviation between DMC and normalized calorimetric profile increases in regions of narrow peaks distant from the melting temperature.  相似文献   
107.
The prevailing view at present is that postsynaptic expression of the classical NMDA receptor-dependent long-term potentiation relies on an increase in the numbers of local AMPA receptors (AMPARs). This is thought to parallel an expansion of postsynaptic cell specializations, for instance dendritic spine heads, which accommodate synaptic receptor proteins. However, glutamate released into the synaptic cleft can normally activate only a hotspot of low-affinity AMPARs that occur in the vicinity of the release site. How the enlargement of the AMPAR pool is causally related to the potentiated AMPAR current remains therefore poorly understood. To understand possible scenarios of postsynaptic potentiation, here we explore a detailed Monte Carlo model of the typical small excitatory synapse. Simulations suggest that approximately 50% increase in the synaptic AMPAR current could be provided by expanding the existing AMPAR pool at the expense of 100–200% new AMPARs added at the same packing density. Alternatively, reducing the inter-receptor distances by only 30–35% could achieve a similar level of current potentiation without any changes in the receptor numbers. The NMDA receptor current also appears sensitive to the NMDA receptor crowding. Our observations provide a quantitative framework for understanding the ‘resource-efficient’ ways to enact use-dependent changes in the architecture of central synapses.  相似文献   
108.
Hypoxia is one of the features of poorly vascularised areas of solid tumours but cancer cells can survive in these areas despite the low oxygen tension. The adaptation to hypoxia requires both biochemical and genetic responses that culminate in a metabolic rearrangement to counter-balance the decrease in energy supply from mitochondrial respiration. The understanding of metabolic adaptations under hypoxia could reveal novel pathways that, if targeted, would lead to specific death of hypoxic regions. In this study, we developed biochemical and metabolomic analyses to assess the effects of hypoxia on cellular metabolism of HCT116 cancer cell line. We utilized an oxygen fluorescent probe in anaerobic cuvettes to study oxygen consumption rates under hypoxic conditions without the need to re-oxygenate the cells and demonstrated that hypoxic cells can maintain active, though diminished, oxidative phosphorylation even at 1% oxygen. These results were further supported by in situ microscopy analysis of mitochondrial NADH oxidation under hypoxia. We then used metabolomic methodologies, utilizing liquid chromatography-mass spectrometry (LC-MS), to determine the metabolic profile of hypoxic cells. This approach revealed the importance of synchronized and regulated catabolism as a mechanism of adaptation to bioenergetic stress. We then confirmed the presence of autophagy under hypoxic conditions and demonstrated that the inhibition of this catabolic process dramatically reduced the ATP levels in hypoxic cells and stimulated hypoxia-induced cell death. These results suggest that under hypoxia, autophagy is required to support ATP production, in addition to glycolysis, and that the inhibition of autophagy might be used to selectively target hypoxic regions of tumours, the most notoriously resistant areas of solid tumours.  相似文献   
109.
Methionine can be reversibly oxidized to methionine sulfoxide (MetO) under physiological conditions. Organisms evolved two distinct methionine sulfoxide reductase families (MSRA & MSRB) to repair oxidized methionine residues. We found that 5 MSRB genes exist in the soybean genome, including GmMSRB1 and two segmentally duplicated gene pairs (GmMSRB2 and GmMSRB5, GmMSRB3 and GmMSRB4). GmMSRB2 and GmMSRB4 proteins showed MSRB activity toward protein-based MetO with either DTT or thioredoxin (TRX) as reductants, whereas GmMSRB1 was active only with DTT. GmMSRB2 had a typical MSRB mechanism with Cys121 and Cys 68 as catalytic and resolving residues, respectively. Surprisingly, this enzyme also possessed the MSRB activity toward free Met-R-O with kinetic parameters similar to those reported for fRMSR from Escherichia coli, an enzyme specific for free Met-R-O. Overexpression of GmMSRB2 or GmMSRB4 in the yeast cytosol supported the growth of the triple MSRA/MSRB/fRMSR (Δ3MSRs) mutant on MetO and protected cells against H2O2-induced stress. Taken together, our data reveal an unexpected diversity of MSRBs in plants and indicate that, in contrast to mammals that cannot reduce free Met-R-O and microorganisms that use fRMSR for this purpose, plants evolved MSRBs for the reduction of both free and protein-based MetO.  相似文献   
110.
Proteins in polyacrylamide gels can be rapidly visualized by soaking in trichloroacetic acid or chloroform followed by illumination with UV light. The UV-light-driven reaction of tryptophan in the presence of trichlorocompounds yields products that emit sufficiently in the visible region to identify the location of the protein bands on the gel. This method can be used to rapidly identify protein bands on a gel in less than 20 min. On thin polyacrylamide gels, 1.0 microg of protein can easily be detected for proteins with typical tryptophan percentages.  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