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   109篇
  免费   8篇
  2018年   3篇
  2017年   3篇
  2016年   2篇
  2015年   2篇
  2014年   5篇
  2013年   8篇
  2012年   3篇
  2011年   5篇
  2010年   6篇
  2009年   3篇
  2008年   5篇
  2007年   6篇
  2006年   5篇
  2005年   7篇
  2004年   1篇
  2003年   1篇
  2002年   3篇
  2001年   5篇
  2000年   8篇
  1999年   6篇
  1997年   2篇
  1996年   2篇
  1992年   2篇
  1991年   1篇
  1990年   3篇
  1989年   2篇
  1988年   3篇
  1987年   3篇
  1986年   1篇
  1985年   2篇
  1978年   1篇
  1976年   2篇
  1975年   1篇
  1974年   1篇
  1971年   1篇
  1967年   1篇
  1966年   2篇
排序方式: 共有117条查询结果,搜索用时 46 毫秒
31.
A short and efficient procedure consisting of two chromatographic steps is described for the isolation of tyrosinase from Neurospora. The first step, Celite-column chromatography, resulted in the isolation of four copper-containing proteins from the crude mycelial extract. Anion-exchange chromatography on DEAE-Sephadex of these proteins resulted in the isolation of electrophoretically and serologically pure tyrosinase. More than 70% of the initial tyrosinase activity was recovered in the final enzyme preparation, which had a specific activity of 450 units/mg.  相似文献   
32.
Summary The inheritance and linkage relationships of a gene for resistance to Fusarium oxysporum f. sp. lycopersici race 1 were analyzed. An interspecific hybrid between a resistant Lycopersicon pennellii and a susceptible L. esculentum was backcrossed to L. esculentum. The genotype of each backcross-1 (BC1) plant with respect to its Fusarium response was determined by means of backcross-2 progeny tests. Resistance was controlled by a single dominant gene, I1, which was not allelic to I, the traditional gene for resistance against the same fungal pathogen that was derived from L. pimpinellifolium. Linkage analysis of 154 molecular markers that segregated in the BC1 population placed I1 between the RFLP markers TG20 and TG128 on chromosome 7. The flanking markers were used to verify the assignment of the I1 genotype in the segregating population. The results are discussed with reference to the possibility of cloning Fusarium resistance genes in tomato.  相似文献   
33.
Phospholipase Cγ isozymes (PLCγ1 and PLCγ2) have a crucial role in the regulation of a variety of cellular functions. Both enzymes have also been implicated in signaling events underlying aberrant cellular responses. Using N-ethyl-N-nitrosourea (ENU) mutagenesis, we have recently identified single point mutations in murine PLCγ2 that lead to spontaneous inflammation and autoimmunity. Here we describe further, mechanistic characterization of two gain-of-function mutations, D993G and Y495C, designated as ALI5 and ALI14. The residue Asp-993, mutated in ALI5, is a conserved residue in the catalytic domain of PLC enzymes. Analysis of PLCγ1 and PLCγ2 with point mutations of this residue showed that removal of the negative charge enhanced PLC activity in response to EGF stimulation or activation by Rac. Measurements of PLC activity in vitro and analysis of membrane binding have suggested that ALI5-type mutations facilitate membrane interactions without compromising substrate binding and hydrolysis. The residue mutated in ALI14 (Tyr-495) is within the spPH domain. Replacement of this residue had no effect on folding of the domain and enhanced Rac activation of PLCγ2 without increasing Rac binding. Importantly, the activation of the ALI14-PLCγ2 and corresponding PLCγ1 variants was enhanced in response to EGF stimulation and bypassed the requirement for phosphorylation of critical tyrosine residues. ALI5- and ALI14-type mutations affected basal activity only slightly; however, their combination resulted in a constitutively active PLC. Based on these data, we suggest that each mutation could compromise auto-inhibition in the inactive PLC, facilitating the activation process; in addition, ALI5-type mutations could enhance membrane interaction in the activated state.Phosphoinositide-specific phospholipase C (PLC)2 enzymes, comprising several families (PLCβ, γ, δ, ϵ, η, and ζ), have been established as crucial signaling molecules involved in regulation of a variety of cellular functions (14). PLC-catalyzed formation of the second messengers, inositol 1,4,5-trisphosphate (IP3) and diacylglycerol, from phosphatidylinositol 4,5-bisphosphate (PIP2), constitutes one of the major cell signaling responses. These second messengers provide a common link from highly specific receptors for hormones, neurotransmitters, antigens, and growth factors to downstream, intracellular targets; thus, they contribute to regulation of biological functions as diverse as cell motility, fertilization, and sensory transduction. Despite this central role for PLC enzymes in signaling networks, the molecular details of their regulation and possible subversion of these regulatory mechanisms in disease remain poorly understood.Of two PLCγ enzymes, PLCγ1 is ubiquitously expressed and appears to regulate a multitude of cellular functions in many tissues. Plcg1-null mice die by embryonic day 9, highlighting the widespread importance of this enzyme (5). PLCγ1 is activated in response to growth factor stimulation; in addition, its function in T-cell responses has been extensively documented (1). PLCγ2, in contrast, is most highly expressed in cells of the hematopoietic system and plays a key role in regulation of the immune response. Consistent with this, Plcg2-null mice display defects in the functioning of B cells, platelets, mast cells, and natural killer cells (6).Both PLCγ enzymes have also been implicated in signaling events underlying aberrant cellular responses. PLCγ1 is critically involved in the regulation of cancer cell motility (711) while PLCγ2 has been implicated in deregulation of the immune responses resembling Btk-dependent X-linked agammaglobulinaemia and SLE disease in humans (1214). It has been suggested that, in cancer cells, PLCγ1 could function as a key, rate-limiting, common component involved in cell motility triggered by several growth factors and integrins (7). In some cancer cells, this increased motility could result from deregulation i.e. higher levels of expression of PLCγ1 (15, 16). The possibility that the activity of PLCγ could be up-regulated due to mutation has not yet been fully investigated in cancer. Previous studies of PLCγ2, however, have demonstrated the first gain-of-function mutation in a PLC molecule in the context of an organism, and shown that, in principle, PLC activity can be greatly enhanced by point mutations (13). Furthermore, this work has demonstrated that such a mutation is linked to a dramatic phenotypic disorder. By using a large scale ENU mutagenesis to discover new immune regulators, several mouse strains were generated with spontaneous autoimmune and inflammatory symptoms; two of these strains harbor a mutation in PLCγ2. In addition to the previously described ALI5 mutation (13) the ALI14 mutation has been identified very recently.3 Strikingly, the well-characterized ALI5 phenotype has shown that the mutation affects many cellular functions deregulated in Plcg2-null mice. Notably, while in Plcg2-null mice such responses are lacking, the ALI5 mutation resulted in their enhancement. In particular, further analyses of the ALI5 mutation in the context of signaling in B-cells have demonstrated that calcium responses to the crosslinking of the B-cell receptor were enhanced and prolonged resulting in enhanced deletion of B cells and autoreactivity (13).The domain organization of PLCγ enzymes is characterized by the insertion of a highly structured region (PLCγ-specific array, γSA) between the two halves of the TIM-barrel catalytic domain common to all PLCs. The γSA comprises a split PH (spPH) domain flanking two tandem SH2 domains and a SH3 domain (1). A distinct regulatory feature of PLCγ enzymes is that their activation is linked to an increase in phosphorylation of specific tyrosine residues (most notably within the γSA) by receptor and non-receptor tyrosine kinases (17, 18). Furthermore, multiple protein-protein interactions (mainly mediated by SH2 domains) also contribute to activation and have an important role in localizing PLCγ into protein complexes with different binding partners, depending on cell type and specific cellular compartments. One mode of activation that is specific for the PLCγ2 isozyme is direct binding to and activation by Rac. The interaction involves the spPH domain, and this activation mechanism does not require tyrosine phosphorylation (19, 20). In molecular terms, changes that lead to PLC activation in response to different input signals, or due to point mutations, are not well understood and require further studies.Here we describe further analysis of the two gain-of-function mutations, ALI5 and ALI14, obtained using ENU mutagenesis. These mutations map to different regions in PLCγ2, and we performed detailed analysis of these regions in both PLCγ isozymes. To characterize the molecular mechanism of gain-of-function, we combined studies in vitro and in different cellular signaling contexts. We have found that ALI5- and ALI14-type point mutations lead, by distinct mechanisms, to an enhancement of responses to a variety of input signals while their combination results in a constitutively active PLC enzyme.  相似文献   
34.
A growing body of work implies that links between PLC isoforms, in particular PLC, and small G-proteins from Ras superfamily could be important in regulation of a number of cellular processes. Through successful use of biochemistry and structural biology, several interactions have been characterized providing some ideas about the regulatory mechanisms. A number of signalling pathways have also been suggested that could involve direct interaction of Ras and Rho GTPases with PLC. Importantly, several studies combining cell biology and genetics have provided new insights into functions of PLC and highlighted the importance of this approach to extend further and consolidate currently incomplete picture regarding its roles in development and disease.  相似文献   
35.

Background

Trans fatty acids are produced either by industrial hydrogenation or by biohydrogenation in the rumens of cows and sheep. Industrial trans fatty acids lower high-density lipoprotein (HDL) cholesterol, raise low-density lipoprotein (LDL) cholesterol, and increase the risk of coronary heart disease. The effects of trans fatty acids from ruminants are less clear. We investigated the effect on blood lipids of cis-9, trans-11 conjugated linoleic acid (CLA), a trans fatty acid largely restricted to ruminant fats.

Methodology/Principal Findings

Sixty-one healthy women and men were sequentially fed each of three diets for three weeks, in random order, for a total of nine weeks. Diets were identical except for 7% of energy (approximately 20 g/day), which was provided either by oleic acid, by industrial trans fatty acids, or by a mixture of 80% cis-9, trans-11 and 20% trans-10, cis-12 CLA. After the oleic acid diet, mean (± SD) serum LDL cholesterol was 2.68±0.62 mmol/L compared to 3.00±0.66 mmol/L after industrial trans fatty acids (p<0.001), and 2.92±0.70 mmol/L after CLA (p<0.001). Compared to oleic acid, HDL-cholesterol was 0.05±0.12 mmol/L lower after industrial trans fatty acids (p = 0.001) and 0.06±0.10 mmol/L lower after CLA (p<0.001). The total-to–HDL cholesterol ratio was 11.6% higher after industrial trans fatty acids (p<0.001) and 10.0% higher after CLA (p<0.001) relative to the oleic acid diet.

Conclusions/Significance

High intakes of an 80∶20 mixture of cis-9, trans-11 and trans-10, cis-12 CLA raise the total to HDL cholesterol ratio in healthy volunteers. The effect of CLA may be somewhat less than that of industrial trans fatty acids.

Trial Registration

ClinicalTrials.gov NCT00529828  相似文献   
36.

Background

Trans fatty acids are produced either by industrial hydrogenation or by biohydrogenation in the rumens of cows and sheep. Industrial trans fatty acids lower HDL cholesterol, raise LDL cholesterol, and increase the risk of coronary heart disease. The effects of conjugated linoleic acid and trans fatty acids from ruminant animals are less clear. We reviewed the literature, estimated the effects trans fatty acids from ruminant sources and of conjugated trans linoleic acid (CLA) on blood lipoproteins, and compared these with industrial trans fatty acids.

Methodology/Principal Findings

We searched Medline and scanned reference lists for intervention trials that reported effects of industrial trans fatty acids, ruminant trans fatty acids or conjugated linoleic acid on LDL and HDL cholesterol in humans. The 39 studies that met our criteria provided results of 29 treatments with industrial trans fatty acids, 6 with ruminant trans fatty acids and 17 with CLA. Control treatments differed between studies; to enable comparison between studies we recalculated for each study what the effect of trans fatty acids on lipoprotein would be if they isocalorically replaced cis mono unsaturated fatty acids. In linear regression analysis the plasma LDL to HDL cholesterol ratio increased by 0.055 (95%CI 0.044–0.066) for each % of dietary energy from industrial trans fatty acids replacing cis monounsaturated fatty acids The increase in the LDL to HDL ratio for each % of energy was 0.038 (95%CI 0.012–0.065) for ruminant trans fatty acids, and 0.043 (95% CI 0.012–0.074) for conjugated linoleic acid (p = 0.99 for difference between CLA and industrial trans fatty acids; p = 0.37 for ruminant versus industrial trans fatty acids).

Conclusions/Significance

Published data suggest that all fatty acids with a double bond in the trans configuration raise the ratio of plasma LDL to HDL cholesterol.  相似文献   
37.
The pleckstrin homology (PH) domains of phospholipase C (PLC)-delta1 and a related catalytically inactive protein, p130, both bind inositol phosphates and inositol lipids. The binding to phosphatidylinositol 4,5-bisphosphate [PtdIns(4,5)P2] by PLC-delta1 is proposed to be the critical interaction required for membrane localization to where the substrate resides; it is also required for the Ca(2+)-dependent activation of PLC-delta1 observed in the permeabilized cells. In the proximity of the PH domain, both PLC-delta1 and p130 possess the EF-hand domain, containing classical motifs implicated in calcium binding. Therefore, in the present study we examined whether the binding of the PH domain to PtdIns(4,5)P2 is regulated by changes in free Ca2+ concentration within the physiological range. A Ca2+ dependent increase in the binding to PtdIns(4,5)P2 was observed with a full-length PLC-delta1, while the isolated PH domain did not show any Ca2+ dependence. However, the connection of the EF-hand motifs to the PH domain restored the Ca2+ dependent increase in binding, even in the absence of the C2 domain. The p130 protein showed similar properties to PLC-delta1, and the EF-hand motifs were again required for the PH domain to exhibit a Ca2+ dependent increase in the binding to PtdIns(4,5)P2. The isolated PH domains from several other proteins which have been demonstrated to bind PtdIns(4,5)P2 showed no Ca2+ dependent enhancement of binding. However, when present within a chimera also containing PLC-delta1 EF-hand motifs, the Ca2+ dependent binding was again observed. These results suggest that the binding of Ca2+ to the EF-hand motifs can modulate binding to PtdIns(4,5)P2 mediated by the PH domain.  相似文献   
38.
39.
The sequence similarity with bacterial neutral sphingomyelinase resulted in the isolation of putative mammalian counterparts and, subsequently, identification of similar molecules in a number of other eukaryotic organisms. Based on sequence similarities and previous characterization of the mammalian enzymes, we have chemically modified specific residues and performed site-directed mutagenesis in order to identify critical catalytic residues and determinants for membrane localization. Modification of histidine residues and the substrate protection experiments demonstrated the presence of reactive histidine residues within the active site. Site directed mutagenesis suggested an essential role in catalysis for two histidine residues (His-136 and His-272), which are conserved in all sequences. Mutations of two additional histidines (His-138 and His-151), conserved only in eukaryotes, resulted in reduced neutral sphingomyelinase activity. In addition to sphingomyelin, the enzyme also hydrolyzed lysophosphatidylcholine. Exposure to an oxidizing environment or modification of cysteine residues using several specific compounds also inactivated the enzyme. Site-directed mutagenesis of eight cysteine residues and gel-shift analysis demonstrated that these residues did not participate in the catalytic reaction and suggested the involvement of cysteines in the formation/breakage of disulfide bonds, which could underlie the reversible inactivation by the oxidizing compounds. Cellular localization studies of a series of deletion mutants, expressed as green fluorescent protein fusion proteins, demonstrated that the transmembrane region contains determinants for the endoplasmic reticulum localization.  相似文献   
40.
Anthracnose, or leaf-curl disease of anemone, caused by Colletotrichum sp., has been reported to occur in Australia, western Europe, and Japan. Symptoms include tissue necrosis, corm rot, leaf crinkles, and characteristic spiral twisting of floral peduncles. Three epidemics of the disease have been recorded in Israel: in 1978, in 1990 to 1993, and in 1996 to 1998. We characterized 92 Colletotrichum isolates associated with anthracnose of anemone (Anemone coronaria L.) for vegetative compatibility (72 isolates) and for molecular genotype (92 isolates) and virulence (4 isolates). Eighty-six of the isolates represented the three epidemics in Israel, one isolate was from Australia, and five isolates originated from western Europe. We divided these isolates into three vegetative-compatibility groups (VCGs). One VCG (ANE-A) included all 10 isolates from the first and second epidemics, and 13 of 62 examined isolates from the third epidemic in Israel, along with the isolate from Australia and 4 of 5 isolates from Europe. Another VCG (ANE-F) included most of the examined isolates (49 of the 62) from the third epidemic, as well as Colletotrichum acutatum from strawberry, in Israel. Based on PCR amplification with species-specific primers, all of the anemone isolates were identified as C. acutatum. Anemone and strawberry isolates of the two VCGs were genotypically similar and indistinguishable when compared by arbitrarily primed PCR of genomic DNA. Only isolate NL-12 from The Netherlands, confirmed as C. acutatum but not compatible with either VCG, had a distinct genotype; this isolate represents a third VCG of C. acutatum. Isolates from anemone and strawberry could infect both plant species in artificial inoculations. VCG ANE-F was recovered from natural infections of both anemone and strawberry, but VCG ANE-A was recovered only from anemone. This study of C. acutatum from anemone illustrates the potential of VCG analysis to reveal distinct subspecific groups within a pathogen population which appears to be genotypically homogeneous by molecular assays.  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