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   1415篇
  免费   107篇
  2021年   7篇
  2020年   4篇
  2019年   6篇
  2018年   13篇
  2017年   18篇
  2016年   17篇
  2015年   40篇
  2014年   47篇
  2013年   80篇
  2012年   65篇
  2011年   74篇
  2010年   30篇
  2009年   43篇
  2008年   83篇
  2007年   73篇
  2006年   65篇
  2005年   82篇
  2004年   80篇
  2003年   72篇
  2002年   71篇
  2001年   43篇
  2000年   53篇
  1999年   41篇
  1998年   26篇
  1997年   27篇
  1996年   23篇
  1995年   33篇
  1994年   15篇
  1993年   8篇
  1992年   22篇
  1991年   23篇
  1990年   26篇
  1989年   18篇
  1988年   20篇
  1987年   19篇
  1986年   12篇
  1985年   17篇
  1984年   18篇
  1983年   7篇
  1982年   7篇
  1981年   6篇
  1979年   5篇
  1978年   16篇
  1976年   8篇
  1975年   4篇
  1974年   9篇
  1973年   5篇
  1972年   7篇
  1963年   3篇
  1958年   3篇
排序方式: 共有1522条查询结果,搜索用时 31 毫秒
951.
952.
The mutagenicity of an oxidized form of dGTP, 8-hydroxy-2′-deoxyguanosine 5′-triphosphate (8-OH-dGTP), was examined using human 293T cells. Shuttle plasmid DNA containing the supF gene was first transfected into the cells, and then 8-OH-dGTP was introduced by means of osmotic pressure. The DNAs replicated in the cells were recovered and then transfected into Escherichia coli. 8-OH-dGTP induced A:T  C:G substitution mutations in the cells. The knock-downs of DNA polymerases η and ζ, and REV1 by siRNAs reduced the A:T  C:G substitution mutations, suggesting that these DNA polymerases are involved in the misincorporation of 8-OH-dGTP opposite A in human cells. In contrast, the knock-down of DNA polymerase ι did not affect the 8-OH-dGTP-induced mutations. The decrease in the induced mutation frequency was more evident by double knock-downs of DNA pols η plus ζ and REV1 plus DNA pol ζ (but not by that of DNA pol η plus REV1), suggesting that REV1-DNA pol η and DNA pol ζ work in different steps. These results indicate that specialized DNA polymerases are involved in the mutagenesis induced by the oxidized dGTP.  相似文献   
953.
Soybean dwarf virus (SbDV), a member of the Luteoviridae family, causes serious yield losses in soybean production in northern Japan. We previously found that an Indonesian soybean cv. Wilis had a high level of resistance to SbDV. Although Wilis is infected by SbDV, symptoms are always mild and develop considerably later compared with many susceptible cultivars. To identify the resistance gene(s) to SbDV in Wilis in a quantitative trait loci (QTL) analysis, we used 71 recombinant inbred lines derived from a cross between Wilis and the susceptible Japanese cv. Toyokomachi and a set of published simple sequence repeat (SSR) markers. The SbDV resistance in Wilis was mainly controlled by a single QTL located near the SSR marker Sat_271 on the linkage group A1. This QTL accounted for 79% of the phenotypic variance. A. Uchibori and J. Sasaki equally contributed to this work.  相似文献   
954.
Our goal is to understand the assembly and regulation of flagellar dyneins, particularly the Chlamydomonas inner arm dynein called I1 dynein. Here, we focus on the uncharacterized I1-dynein IC IC97. The IC97 gene encodes a novel IC without notable structural domains. IC97 shares homology with the murine lung adenoma susceptibility 1 (Las1) protein—a candidate tumor suppressor gene implicated in lung tumorigenesis. Multiple, independent biochemical assays determined that IC97 interacts with both α- and β-tubulin subunits within the axoneme. I1-dynein assembly mutants suggest that IC97 interacts with both the IC138 and IC140 subunits within the I1-dynein motor complex and that IC97 is part of a regulatory complex that contains IC138. Microtubule sliding assays, using axonemes containing I1 dynein but devoid of IC97, show reduced microtubule sliding velocities that are not rescued by kinase inhibitors, revealing a critical role for IC97 in I1-dynein function and control of dynein-driven motility.  相似文献   
955.
956.
Growth-blocking peptide (GBP) is a hormone-like peptide that suppresses the growth of the host armyworm. Although the 23-amino acid GBP (1–23 GBP) is expressed in nonparasitized armyworm plasma, the parasitization by wasp produces the 28-amino acid GBP (1–28 GBP) through an elongation of the C-terminal amino acid sequence. In this study, we characterized the GBP variants, which consist of various lengths of the C-terminal region, by comparing their biological activities and three-dimensional structures. The results of an injection study indicate that 1–28 GBP most strongly suppresses larval growth. NMR analysis shows that these peptides have basically the same tertiary structures and that the extension of the C-terminal region is disordered. However, the C-terminal region of 1–28 GBP undergoes a conformational transition from a random coiled state to an α-helical state in the presence of dodecylphosphocholine micelles. This suggests that binding of the C-terminal region would affect larval growth activity.Growth-blocking peptide (GBP)2 was initially identified from the hemolymph of armyworm Pseudaletia separata as a 25-amino acid peptide (1–25 GBP) that prevents the onset of pupation of the host by parasitization of wasp Cotesia kariyai (14). Injection of GBP into nonparasitized armyworm larvae early in the last instar delays larval growth and retards pupation for more than a few days. Our previous studies showed that GBP is a hormone-like biogenic peptide of the host armyworm (5, 6). In nonparasitized larvae, the concentrations of GBP were much higher in the early larval stages than in the latter ones. However, parasitization by wasp induces an elevation of GBP in the last larval stages. This elevation was shown to lead to growth retardation via repression of juvenile hormone esterase activity (79). Interestingly, a cDNA analysis indicated that the cDNA encodes a 23-amino acid GBP (1–23 GBP), although GBP purified from parasitized armyworm plasma consists of 25 amino acid residues. GBP was expressed as a 23-residue peptide (1–23 GBP) in nonparasitized armyworm larvae, whereas 1–25 GBP, containing Tyr24 and Gln25, was purified from the parasitized larvae. Moreover, the preliminary peptide sequencing of GBP prepared from parasitized larval hemolymph showed the 26th and 27th residues on rare occasions (Leu and Ile, respectively) (6). On the basis of these results, we concluded that the TAG stop codon for the 24th amino acid was unusually decoded as Tyr upon parasitization by parasitoid wasps (10) and predicted that an intact and mature GBP synthesized in the parasitized armyworm larvae would consist of 28 amino acid residues (1–28 GBP).GBP has multiple functions: adhesion and spreading of a specific class of immune cells (plasmatocytes), proliferation of various cultured cells, and induction of larval paralysis (1113). More than 10 GBP homologous peptides have been identified in Lepidopteran insects, and based on their N-terminal consensus sequences (Glu1-Asn2-Phe3), they have been categorized as the ENF peptide family (14). The tertiary structure of 1–25 GBP consists of a disordered N-terminal region (residues Glu1–Gly6), a well ordered core region (residues Cys7–Thr22) stabilized by a disulfide bond and a short antiparallel β-sheet, and a short unstructured C-terminal region (Phe23–Glu25) (15). Because no GBP receptor or its gene has been isolated yet, the nature of either of them at the cellular and molecular levels is poorly understood at present. In contrast, the relationship between the structure and activity of GBP has been well studied by analyzing the biological activities of several variants of GBP and plasmatocyte-spreading peptide (one of the ENF family peptides). Especially, extensive studies on the N termini (residues 1–6) of GBP and plasmatocyte-spreading peptide demonstrated the importance of Phe3 for exerting their hemocyte stimulating activity, thereby suggesting a possible mechanism for receptor activation that requires binding of the aromatic ring of Phe3 and a closely spaced primary amine with receptor activating properties (1619).In contrast, the C termini of GBP and other ENF peptides have received less attention, because of the weak secondary structure predictions. Therefore, in this study we focused on the C terminus region of GBP and analyzed its contribution to the expression of some biological activities and to the tertiary structure of this peptide. Especially, we prepared GBP with 28 amino acids and characterized the C-terminal region of 1–28 GBP (residues Phe23–Thr28), because we knew that GBP is present as a 23-amino acid peptide in nonparasitized healthy larvae and that GBP with 28 amino acids has been found only in parasitized host larvae. Our results suggest that the elongation of the C-terminal region of Phe23–Thr28 greatly reinforced GBP binding with the membrane. Further, the elongation increased GBP inhibition of larval growth.  相似文献   
957.
Hideyuki Adachi  Isao Enami  Nobuo Kamiya 《BBA》2009,1787(2):121-128
Crystal structure of photosystem II (PSII) has been reported from prokaryotic cyanobacteria but not from any eukaryotes. In the present study, we improved the purification procedure of PSII dimers from an acidophilic, thermophilic red alga Cyanidium caldarium, and crystallized them in two forms under different crystallization conditions. One had a space group of P2221 with unit cell constants of a = 146.8 Å, b = 176.9 Å, and c = 353.7 Å, and the other one had a space group of P212121 with unit cell constants of a = 209.2 Å, b = 237.5 Å, and c = 299.8 Å. The unit cell constants of both crystals and the space group of the first-type crystals are different from those of cyanobacterial crystals, which may reflect the structural differences between the red algal and cyanobacterial PSII, as the former contains a fourth extrinsic protein of 20 kDa. X-ray diffraction data were collected and processed to a 3.8 Å resolution with the first type crystal. For the second type crystal, a post-crystallization treatment of dehydration was employed to improve the resolution, resulting in a diffraction data of 3.5 Å resolution. Analysis of this type of crystal revealed that there are 2 PSII dimers in each asymmetric unit, giving rise to 16 PSII monomers in each unit cell, which contrasts to 4 dimers per unit cell in cyanobacterial crystals. The molecular packing of PSII within the unit cell was constructed with the molecular replacement method and compared with that of the cyanobacterial crystals.  相似文献   
958.
959.
A cytochrome P450cam monooxygenase (P450cam) system from the soil bacterium Pseudomonas putida requires electron transfer among three different proteins and a cofactor, nicotinamide adenine dinucleotide (NADH), for oxygenation of its natural substrate, camphor. Herein, we report a facile way to significantly enhance the catalytic efficiency of the P450cam system by the coupling of its native electron transfer system with enzymatic NADH regeneration catalyzed by glycerol dehydrogenase (GLD) in Escherichia coli whole cell biocatalysts. Recombinant E. coli harboring the P450cam system, but lacking GLD, exhibited little activity for camphor hydroxylation. In contrast, coexpression of GLD with the proteinaceous electron transfer components of P450cam resulted in about tenfold improvement in the substrate conversion, implying that the whole cell biocatalyst utilized molecular oxygen, endogenous NADH, and glycerol in the cell for catalysis. The addition of glycerol to the reaction media further promoted camphor hydroxylation, suggesting that exogenous glycerol is also available for GLD in the host cell and actively participates in the catalytic cycle. These results clearly show the utility of GLD towards functional reconstruction of the native P450cam system. The present approach may also be useful for E. coli whole cell biocatalysts with the other NADH-dependent oxygenases and oxidoreductases.  相似文献   
960.
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