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   15569篇
  免费   1534篇
  国内免费   2篇
  2024年   16篇
  2023年   97篇
  2022年   216篇
  2021年   375篇
  2020年   233篇
  2019年   272篇
  2018年   324篇
  2017年   294篇
  2016年   506篇
  2015年   954篇
  2014年   923篇
  2013年   1127篇
  2012年   1468篇
  2011年   1403篇
  2010年   894篇
  2009年   767篇
  2008年   995篇
  2007年   993篇
  2006年   946篇
  2005年   903篇
  2004年   859篇
  2003年   737篇
  2002年   679篇
  2001年   112篇
  2000年   69篇
  1999年   113篇
  1998年   135篇
  1997年   76篇
  1996年   71篇
  1995年   53篇
  1994年   54篇
  1993年   52篇
  1992年   36篇
  1991年   45篇
  1990年   30篇
  1989年   16篇
  1988年   31篇
  1987年   15篇
  1986年   16篇
  1985年   18篇
  1984年   14篇
  1983年   17篇
  1982年   20篇
  1981年   12篇
  1980年   16篇
  1979年   7篇
  1978年   11篇
  1977年   14篇
  1976年   7篇
  1974年   11篇
排序方式: 共有10000条查询结果,搜索用时 24 毫秒
991.
In this work we describe a simple two step separation procedure for the separation and purification of short DNA fragments. The first step involves precipitating the DNA using the cationic surfactant dodecyltrimethylammonium bromide. Dodecyltrimethylammonium bromide, unlike cetyltrimethylammonium bromide will not precipitate DNA before complexation is complete thus providing a high purity DNA. The second step involves dissolution of the DNA-dodecyltrimethylammonium complex in 75% ethanol, followed by precipitation of the Sodium-DNA salt, by titrating in a salt solution. This method is particularly suited to purification of short fragments as it does not require high salt concentrations in the ethanol precipitation step, which can be damaging for short DNA. The ability of dodecyltrimethylammonium bromide to remove ethidium bromide from intercalation sites on the DNA is also discussed  相似文献   
992.
Rat Chromosome 10 (RNO10) harbors Cia5, a non-MHC quantitative trait locus (QTL) that regulates the severity of type II collagen-induced arthritis (CIA) in DAxF344 and DAxBN F2 rats. CIA is an animal model with many features that resemble rheumatoid arthritis. To facilitate analysis of Cia5 independently of the other CIA regulatory loci on other chromosomes, DA recombinant QTL speed congenic rats, DA.F344(Cia5), were generated. These QTL congenic rats have a large chromosomal segment containing Cia5 (interval size < or =80.1 cM) from CIA-resistant F344 rats introgressed into their genome. Phenotypic analyses of these rats for susceptibility and severity of CIA confirmed that Cia5 is an important disease-modifying locus. CIA severity was significantly lower in the Cia5 congenic rats than in DA controls. We also generated DA Cia5 speed sub-congenic rats, DA.F344(Cia5a), which had a smaller segment of the F344 genome, Cia5a, comprising only the distal q-telomeric end (interval size < or = 22.5 cM) of Cia5, introgressed into their genome. DA.F344(Cia5a) sub-congenic rats also exhibited reduced CIA disease severity compared with the parental DA rats. The regulatory effects in both congenic strains were sex influenced. The disease-ameliorating effect of the larger fragment, Cia5, was greater in males than in females, but the effect of the smaller fragment, Cia5a, was greater in females. We also present an improved genetic linkage map covering the Cia5/Cia5a region, which we have integrated with two rat radiation hybrid maps. Comparative homology analysis of this genomic region with mouse and human chromosomes was also undertaken. Regulatory loci for multiple autoimmune/inflammatory diseases in rats (RNO10), mice (MMU11), and humans (HSA17 and HSA5q23-q31) map to chromosomal segments homologous to Cia5 and Cia5a.  相似文献   
993.
994.
995.
Womack JS  Randall J  Kemp JD 《Planta》2000,210(5):844-847
 A previously unidentified extension of an open reading frame from the genomic DNA of Japonica rice (Oryza sativa L.) encoding oryzacystatin-I (OC-I; access. M29259, protein ID AAA33912.1) has been identified as a 5′ gene segment coding for the OC-I signal peptide. The signal peptide appears to direct a pre-protein (SPOC-I; Accession No. AF164378) to the endoplasmic reticulum, where it is processed into the mature form of OC-I. The start codon of SPOC-I begins 114 bp upstream from that previously published for OC-I. A putative proteolytic site, which may yield a mature OC-I approximately 12 residues larger than previously described, has been identified within SPOC-I between Ala-26 and Glu-27. The signal peptide sequence was amplified by polymerase chain reaction using genomic DNA from O. sativa seedlings and ligated to the 5′ end of the truncated OC-I gene at the endogenous SalI site. Partially purified protein extracts from Escherichia coli expressing SPOC-I reacted with polyclonal antibodies raised against OC-I and revealed a protein of the expected molecular weight (15,355 Da). In-vitro translation of SPOC-I in the presence of microsomal membranes yielded a processed product approximately 2.7 kDa smaller than the pre-protein. Nicotiana tabacum L. cv. Xanthi plants independently transformed with the SPOC-I gene processed SPOC-I and accumulated the mature form of OC-I (approximately 12.6 kDa), which co-migrated with natural, mature OC-I extracted from rice seed when separated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Received: 29 July 1999 / Accepted: 25 August 1999  相似文献   
996.
Ozone: a tool for probing programmed cell death in plants   总被引:27,自引:0,他引:27  
  相似文献   
997.
Pseudomonas strain ADP metabolizes the herbicide atrazine via three enzymatic steps, encoded by the genes atzABC, to yield cyanuric acid, a nitrogen source for many bacteria. Here, we show that five geographically distinct atrazine-degrading bacteria contain genes homologous to atzA, -B, and -C. The sequence identities of the atz genes from different atrazine-degrading bacteria were greater than 99% in all pairwise comparisons. This differs from bacterial genes involved in the catabolism of other chlorinated compounds, for which the average sequence identity in pairwise comparisons of the known members of a class ranged from 25 to 56%. Our results indicate that globally distributed atrazine-catabolic genes are highly conserved in diverse genera of bacteria.Atrazine [2-chloro-4-(ethylamino)-6-(isopropylamino)- 1,3,5-triazine] is a herbicide used for controlling broad-leaf and grassy weeds and is relatively persistent in soils (51). Atrazine and other s-triazine compounds have been detected in ground and surface waters at levels exceeding the Environmental Protection Agency’s maximum contaminant level of 3 ppb (30).Microbial populations exposed to synthetic chlorinated compounds, such as atrazine, often respond by producing enzymes that degrade these molecules. Most of our current understanding of the genes and enzymes involved in atrazine degradation derives from studies using Pseudomonas strain ADP, in which the first three enzymatic steps in atrazine degradation have been defined (6, 14, 15, 48). The genes atz A, -B, and -C, which encode these enzymes, have been cloned and sequenced. Atrazine chlorohydrolase (AtzA), hydroxyatrazine ethylaminohydrolase (AtzB), and N-isopropylammelide isopropylaminohydrolase (AtzC) sequentially convert atrazine to cyanuric acid (6, 14, 15, 48) (Fig. (Fig.1).1). Cyanuric acid and related compounds are catabolized by many soil bacteria (10, 11, 17, 24, 26, 61), and by Pseudomonas sp. ADP, to carbon dioxide and ammonia (35). This provides the evolutionary pressure for the atzA, -B, and -C genes to permit bacterial growth on the more than one billion pounds of atrazine that have been applied to soils globally (20). Here we used a knowledge of the atzA, -B, and -C gene sequences to investigate the presence of homologous genes in other atrazine-degrading bacteria. In this study, we report that five atrazine-degrading microorganisms, which were recently isolated from geographically separated sites exposed to atrazine, contained nearly identical atzA, -B, and -C genes. Open in a separate windowFIG. 1Pathway for atrazine catabolism to cyanuric acid in Pseudomonas sp. strain ADP.

Atrazine-catabolizing bacteria used in this study.

Until recently, attempts at isolating bacteria (18) or fungi (27) that completely degrade atrazine to carbon dioxide, ammonia, and chloride were unsuccessful. While several microorganisms were shown to dealkylate atrazine, they were unable to displace the chlorine atom (41, 54). Since 1994, several research groups have independently isolated atrazine-degrading bacteria that displaced the chlorine atom and mineralized atrazine (3, 7, 13, 35, 39, 46). Six of these bacterial cultures, listed in Table Table1,1, were studied here, and the Clavibacter strain had been investigated previously (13).

TABLE 1

Recently isolated atrazine-catabolizing bacteria
GenusStrainLocation where isolatedYr reported (reference)
PseudomonasaADPAgricultural-chemical dealership site, Little Falls, Minn.1995 (35)
RalstoniaaM91-3Agricultural soil, Ohio1995 (46, 55)
Mixed cultureBasel, Switzerland1995 (57)
ClavibacterAgricultural soil, Riverside, Calif.1996 (13)
AgrobacteriumJ14aAgricultural soil, Nebraska1996 (39)
NDb38/38Atrazine-contaminated soil, Indiana1996 (3)
AlcaligenesaSG1Industrial settling pond, San Gabriel, La.1997 (7)
Open in a separate windowaIsolate identity based on 16S rRNA sequence analysis. bND, not determined. 

Detection of atzA, -B, and -C homologs in atrazine-degrading microorganisms by PCR analysis.

Recently isolated atrazine-degrading bacteria were screened for the presence of DNA homologous to the Pseudomonas strain ADP atzABC genes, which encode enzymes transforming atrazine to cyanuric acid (Fig. (Fig.1).1). Total genomic DNA was isolated from each of these bacteria as described elsewhere (49), and the PCR technique was used to amplify sequences internal to the atzA, -B, and -C genes as described elsewhere (13). Custom primers were designed specifically for atzA (5′CCATGTGAACCAGATCCT3′ and 5′TGAAGCGTCCACATTACC3′), atzB (5′TCACCGGGGATGTCGCGGGC3′ and 5′CTCTCCCGCATGGCATCGGG3′), and atzC (5′GCTCACATGCAGGTACTCCA3′ and 5′GTACCATATCACCGTTTGCCA3′) by using the Primer Designer package, version 2.01 (Scientific and Educational Software, State Line, Pa.), and were synthesized by Gibco BRL (Gaithersburg, Md.). PCR fragments were amplified by using Taq DNA polymerase (Gibco BRL) (22) and were separated from primers on a 1.0% agarose gel. The results of these studies (Fig. (Fig.2)2) indicated that PCR amplification consistently produced DNA fragments of 0.5 kb for all organisms when the atzA or -B primers were used and fragments of 0.6 kb when the atzC primers were used. Open in a separate windowFIG. 2PCR analysis with primers designed to amplify internal regions of atzA (lanes 1 to 5), atzB (lanes 6 to 10), and atzC (lanes 11 to 15). The atrazine-degrading bacteria analyzed were Pseudomonas strain ADP (35) (lanes 1, 6, and 11), Alcaligenes strain SGI (7) (lanes 2, 7, and 12), Ralstonia strain M91-3 (46) (lanes 3, 8, and 13), Agrobacterium strain J14a (39) (lanes 4, 9, and 14), and isolate 38/38 (3) (lanes 5, 10, and 15). Values to the right of the gel are sizes (in kilobase pairs).Southern hybridization analyses were performed on the PCR-amplified DNA as described elsewhere (49) to confirm the presence of homologous DNA. We used a 0.6-kb ApaI/PstI fragment from pMD4 (15), a 1.5-kb BglII fragment from pATZB-2 (6), and a 2.0-kb EcoRI/AvaI fragment from pTD2.5 (48) as probes for atzA, -B, and -C genes, respectively. DNA probes were labeled with [α-32P]dCTP by using the Rediprime Random Primer Labeling Kit (Amersham Life Science, Arlington Heights, Ill.) according to the manufacturer’s instructions. Southern hybridization analyses, performed under stringent conditions, confirmed that each strain contained DNA homologous to atzA, -B, and -C (data not shown). With strain M91-3 and isolate 38/38, however, in addition to the expected 0.5-kb atzB PCR product (Fig. (Fig.2,2, lanes 8 and 10), a 1.2-kb fragment was also obtained. However, no hybridization to this fragment was seen with the atzB probe. Similar investigations showed that a mixed culture obtained from Switzerland (Table (Table1),1), capable of degrading atrazine, also contained DNA homologous to all three atz genes (12).As a negative control, bacteria known not to degrade atrazine were analyzed. PCR analyses were carried out with genomic DNA from the following randomly chosen laboratory strains: Rhodococcus chlorophenolicus (1), Flavobacterium sp. (47), Streptomyces coelicolor M145 (21), Amycolatopsis mediterranei (19), Agrobacterium strain A136 and strain A348 (A136/pTiA6NC) (60), Arthrobacter globiformis MN1 (45), Bradyrhizobium japonicum (33), Rhizobium sp. strain NGR 234 (44), Pseudomonas NRRLB12228, and Klebsiella pneumoniae 99 (16). None of these strains contained DNA that was amplified by PCR using the primers designed to identify the atzA, -B, or -C gene (data not shown).

DNA sequences of atzA, -B, and -C homologs in atrazine-degrading microorganisms.

DNAs amplified from the five strains in Table Table11 with the atzA, -B, and -C primers were purified from gel slices by using the GeneClean II System (Bio 101, Inc., Vista, Calif.) and sequenced with a PRISM Ready Reaction DyeDeoxy Terminator Cycle Sequencing kit (Perkin-Elmer Corp., Norwalk, Conn.) and an ABI model 373A DNA sequencer (Applied Biosystems, Foster City, Calif.). The GCG sequence analysis software package (Genetics Computer Group, Inc., Madison, Wis.) was used for all DNA and protein sequence comparisons and alignments. Table Table22 summarizes these data. The PCR-amplified genes were ≥99% identical to the Pseudomonas strain ADP atzA, -B, and -C genes in all pairwise comparisons of DNA sequences. This remarkable sequence identity suggested that each atz gene in the different genera was derived from a common ancestor and that they have diverged evolutionarily only to a limited extent.

TABLE 2

Sequence identities of atzABC homologs from different atrazine-degrading bacteria
Strain% DNA sequence identitya
atzAatzBatzC
Pseudomonas ADP100100100
Alcaligenes SG199.2100100
Ralstonia M91-399.0100100
Agrobacterium J14a99.1100100
Isolate 38/3899.310099.8
Open in a separate windowaDNA sequences obtained from each strain by using the ataA, -B, and -C primers were compared with the atzABC gene sequences from Pseudomonas strain ADP. A review of the literature on other bacterial catabolic pathways indicated a much greater degree of divergence when genes encoding enzymes for the catabolism of other commercially relevant chlorinated compounds were compared (Table (Table3).3). As with atrazine, multiple bacterial strains that catabolize 1,2-dichloroethane, chloroacetic acid, 2,4-dichlorophenoxyacetate, dichloromethane, and 4-chlorobenzoate have been isolated. A comparison of the gene sequences encoding the initiating reactions in the catabolism of each of those compounds revealed that sequence divergence was comparatively high. In pairwise comparisons within each gene class, the average sequence identities ranged from 25 to 56% (divergence was 46 to 75%). With the atzABC genes, by contrast, there is at most a 1% sequence difference within the sequenced gene region (Table (Table2).2). Moreover, the atzB sequences were completely identical, and the atzC genes diverged by only 1 bp in one of the five strains tested. This suggests that the atz genes recently arose from a single origin and have become distributed globally. Similarly, identical parathion hydrolase genes were isolated from two bacteria representing different genera and global locations (40, 52, 53).

TABLE 3

Sequence comparisons of isofunctional bacterial enzymes that catabolize chlorinated compounds
GeneEnzymeAverage % protein sequence identitya (no. of pairwise comparisons)References
dhlA, dhaAHaloalkane dehalogenase25.0 (1)23, 31
dehC, hadL, dehH, dehH1, dehH2, dhlB, dehCI, dehCII2-Haloacid dehalogenase36.6 ± 3.9 (36)5, 25, 28, 29, 42, 43, 50, 59
tfdA2,4-Dichlorophenoxyacetate monooxygenase43.2 ± 4.6 (21)b34, 37, 38, 56, 58
dcmADichloromethane dehalogenase56.0 (1)4, 32
atzAAtrazine chlorohydrolase98.6 ± 0.12 (15)cThis study
atzBHydroxyatrazine ethylaminohydrolase100 (10)cThis study
atzCN-Isopropylammelide isopropylaminohydrolase99.0 ± 0.43 (10)cThis study
Open in a separate windowaAll possible pairwise alignments of translated gene sequences were made. The average percent identity is the mean of the percent identity values for all pairwise alignments ± standard error of the mean. bIncludes full protein sequences as well as partial protein sequences of ≥100 amino acids. cSequence identity within a 0.5-kb PCR product for atzA and -B and within a 0.6-kb PCR product for atzC. Six sequences were analyzed for atzA, and five were analyzed for atzB and -C. The data presented here provide further support for previous studies suggesting that hydroxyatrazine in the environment derives from biological processes (36), and not solely from abiotic reactions (2, 9). The present data, and a recent report by Bouquard et al. (8), indicate that the gene encoding atrazine chlorohydrolase is widespread in the United States and Europe.Our observations argue for a single, recent evolutionary origin of the atz genes and their subsequent global distribution. We have recently localized the atzA, -B, and -C genes to a large, self-transmissible plasmid in Pseudomonas strain ADP (12), and possible mechanisms of transfer of the atzABC genes are currently under investigation.  相似文献   
998.
The 26S proteasome complex plays a major role in the non-lysosomal degradation of intracellular proteins. Purified 26S proteasomes give a pattern of more than 40 spots on 2D-PAGE gels. The positions of subunits have been identified by mass spectrometry of tryptic peptides and by immunoblotting with subunit-specific antipeptide antibodies. Two-dimensional polyacrylamide gel electrophoresis of proteasomes immunoprecipitated from [32P]phosphate-labelled human embryo lung L-132 cells revealed the presence of at least three major phosphorylated polypeptides among the regulatory subunits as well as the C8 and C9 components of the core 20S proteasome. Comparison with the positions of the regulatory polypeptides revealed a minor phosphorylated form to be S7 (MSS1). Antibodies against S4, S6 (TBP7) and S12 (MOV34) all cross-reacted at the position of major phosphorylated polypeptides suggesting that several of the ATPase subunits may be phosphorylated. The phosphorylation of S4 was confirmed by double immunoprecipitation experiments in which 26S proteasomes were immunoprecipitated as above and dissociated and then S4 was immunoprecipitated with subunit-specific antibodies. Antibodies against the non-ATPase subunit S10, which has been suggested by others to be phosphorylated, did not coincide with the position of a phosphorylated polypeptide. Some differences were observed in the 2D-PAGE pattern of proteasomes immunoprecipitated from cultured cells compared to purified rat liver 26S proteasomes suggesting possible differences in subunit compositions of 26S proteasomes.  相似文献   
999.
The yeast YCC5 gene encodes a putative amino acid permease and is homologous to GNP1 (encoding a high-affinity glutamine permease). Using strains with disruptions in the genes for multiple permeases, we demonstrated that Ycc5 (which we have renamed Agp1) is involved in the transport of asparagine and glutamine, performed a kinetic analysis of this activity, and showed that AGP1 expression is subject to nitrogen repression.  相似文献   
1000.
The nonreceptor protein tyrosine kinase ZAP-70 is a critical enzyme required for successful T lymphocyte activation. After antigenic stimulation, ZAP-70 rapidly associates with T cell receptor (TCR) subunits. The kinetics of its translocation to the cell surface, the properties of its specific interaction with the TCRζ chain expressed as a chimeric protein (TTζ and Tζζ), and its mobility in different intracellular compartments were studied in individual live HeLa cells, using ZAP-70 and Tζζ fused to green fluorescent protein (ZAP-70 GFP and Tζζ–GFP, respectively). Time-lapse imaging using confocal microscopy indicated that the activation-induced redistribution of ZAP-70 to the plasma membrane, after a delayed onset, is of long duration. The presence of the TCRζ chain is critical for the redistribution, which is enhanced when an active form of the protein tyrosine kinase Lck is coexpressed. Binding specificity to TTζ was indicated using mutant ZAP-70 GFPs and a truncated ζ chimera. Photobleaching techniques revealed that ZAP-70 GFP has decreased mobility at the plasma membrane, in contrast to its rapid mobility in the cytosol and nucleus. Tζζ– GFP is relatively immobile, while peripherally located ZAP-70 in stimulated cells is less mobile than cytosolic ZAP-70 in unstimulated cells, a phenotype confirmed by determining the respective diffusion constants. Examination of the specific molecular association of signaling proteins using these approaches has provided new insights into the TCRζ–ZAP-70 interaction and will be a powerful tool for continuing studies of lymphocyte activation.  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