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   167篇
  免费   3篇
  170篇
  2023年   4篇
  2022年   2篇
  2021年   8篇
  2020年   2篇
  2019年   3篇
  2018年   5篇
  2017年   2篇
  2016年   7篇
  2015年   5篇
  2014年   12篇
  2013年   8篇
  2012年   14篇
  2011年   14篇
  2010年   11篇
  2009年   15篇
  2008年   15篇
  2007年   10篇
  2006年   10篇
  2005年   2篇
  2004年   4篇
  2003年   8篇
  2002年   4篇
  2000年   1篇
  1999年   3篇
  1998年   1篇
排序方式: 共有170条查询结果,搜索用时 15 毫秒
61.
62.

Background  

We have recently reported the expression of murine Implantation Serine Proteinase genes in pre-implantation embryos (ISP1) and uterus (ISP1 and ISP2). These proteinases belong to the S1 proteinase family and are similar to mast cell tryptases, which function as multimers.  相似文献   
63.
The GroEL/GroES protein folding chamber is formed and dissociated by ATP binding and hydrolysis. ATP hydrolysis in the GroES-bound (cis) ring gates entry of ATP into the opposite unoccupied trans ring, which allosterically ejects cis ligands. While earlier studies suggested that hydrolysis of cis ATP is the rate-limiting step of the cycle (t½ ∼ 10 s), a recent study suggested that ADP release from the cis ring may be rate-limiting (t½ ∼ 15-20 s). Here we have measured ADP release using a coupled enzyme assay and observed a t½ for release of ?4-5 s, indicating that this is not the rate-limiting step of the reaction cycle.  相似文献   
64.
65.
Thyrotropin-releasing hormone (TRH) analogues in which the N(1)-position of the imidazole ring of the centrally placed histidine residue is substituted with various alkyl groups were synthesized and studied as agonists for TRH receptor subtype 1 (TRH-R1) and subtype 2 (TRH-R2). Analogue 3 (R=C2H5) exhibited binding affinity (Ki) of 0.012 microM to TRH-R1 that is about 1.1-fold higher than that of TRH. Several analogues were found to selectively activate TRH-R2 with greater potency than TRH-R1. The most selective agonist of the series 5 [R=CH(CH3)2] was found to activate TRH-R2 with a potency (EC50) of 0.018 microM but could only activate TRH-R1 at EC50 value of 1.6 microM; that is, exhibited 88-fold greater potency for TRH-R2 versus TRH-R1. The results of this study indicate that modulation of central histidine residue is important for designing analogues which were selective agonist at TRH receptor subtypes.  相似文献   
66.
Plant peroxisomes are highly dynamic organelles that mediate a suite of metabolic processes crucial to development. Peroxisomes in seeds/dark-grown seedlings and in photosynthetic tissues constitute two major subtypes of plant peroxisomes, which had been postulated to contain distinct primary biochemical properties. Multiple in-depth proteomic analyses had been performed on leaf peroxisomes, yet the major makeup of peroxisomes in seeds or dark-grown seedlings remained unclear. To compare the metabolic pathways of the two dominant plant peroxisomal subtypes and discover new peroxisomal proteins that function specifically during seed germination, we performed proteomic analysis of peroxisomes from etiolated Arabidopsis (Arabidopsis thaliana) seedlings. The detection of 77 peroxisomal proteins allowed us to perform comparative analysis with the peroxisomal proteome of green leaves, which revealed a large overlap between these two primary peroxisomal variants. Subcellular targeting analysis by fluorescence microscopy validated around 10 new peroxisomal proteins in Arabidopsis. Mutant analysis suggested the role of the cysteine protease RESPONSE TO DROUGHT21A-LIKE1 in β-oxidation, seed germination, and growth. This work provides a much-needed road map of a major type of plant peroxisome and has established a basis for future investigations of peroxisomal proteolytic processes to understand their roles in development and in plant interaction with the environment.Peroxisomes, originally known as microbodies, are small and single-membrane eukaryotic organelles that compartmentalize various oxidative metabolic functions. Most peroxisomal matrix proteins carry a C-terminal tripeptide named PEROXISOME TARGETING SIGNAL TYPE1 (PTS1), and fewer contain an N-terminal nonapeptide, PTS2 (Lanyon-Hogg et al., 2010). PTS1 is further divided into major and minor PTS1s. Major PTS1 tripeptides, such as SKL> and SRL> (> represents the stop codon), are by themselves sufficient to direct a protein to the peroxisome (Reumann, 2004), whereas minor PTS1s are usually found in low-abundance proteins and require additional upstream elements for peroxisomal targeting (Kaur et al., 2009). Peroxisomes are highly variable morphologically and metabolically, as their size, shape, abundance, and enzymatic content can differ depending on the species, tissue and cell type, and prevailing environmental conditions (Beevers, 1979; van den Bosch et al., 1992; Kaur et al., 2009; Hu et al., 2012; Schrader et al., 2012).Plant peroxisomes participate in a wide range of metabolic processes, such as lipid metabolism, photorespiration, detoxification, biosynthesis of jasmonic acid, and metabolism of indole-3-butyric acid (IBA), nitrogen, sulfite, and polyamine (Kaur et al., 2009; Hu et al., 2012). Specific names had been given to certain types of peroxisomes due to their unique metabolic properties. For example, the term glyoxysome was coined when a new type of organelle that contained enzymes of the glyoxylate cycle was identified from the endosperm of castor bean (Ricinus communis; Breidenbach et al., 1968). It was later realized that glyoxysomes are in fact a type of peroxisome, and Beevers (1979) subsequently classified plant peroxisomes into three subtypes based on their primary biochemical functions. Glyoxysomes are located in storage organs such as fatty seedling tissues and play a major role in converting fatty acids to sugar; leaf peroxisomes are involved in photorespiration; and nonspecialized peroxisomes exist in other plant tissues and perform unknown functions.The primary function of leaf peroxisomes is the recycling of phosphoglycolate during photorespiration, a process coordinated by chloroplasts, peroxisomes, mitochondria, and the cytosol. In this pathway, phosphoglycolate produced by the oxygenase activity of Rubisco is ultimately converted to glycerate, which reenters the chloroplastic Calvin-Benson cycle (Foyer et al., 2009; Peterhansel et al., 2010). The peroxisome-localized enzymes glycolate oxidase (GOX), catalase, aminotransferase (serine:glyoxylate aminotransferase [SGT] and glutamate-glyoxylate aminotransferase [GGT]), HYDROXYPYRUVATE REDUCTASE1 (HPR1), and peroxisomal malate dehydrogenase (PMDH) are involved in the process (Reumann and Weber, 2006). On the other hand, lipid mobilization through fatty acid β-oxidation and the glyoxylate cycle is the main function for peroxisomes in seeds and germinating seedlings. In this process, fatty acids are first activated into fatty acyl-CoA esters by the acyl-activating enzyme (AAE)/acyl-CoA synthetase before entering the β-oxidation cycle, during which an acetyl-CoA is cleaved in each cycle by the successive action of acyl-CoA oxidase (ACX), multifunctional protein (MFP), and 3-keto-acyl-CoA thiolase (KAT). Acetyl-CoA, an end product of β-oxidation, is further converted to four-carbon carbohydrates by the glyoxylate cycle, in which isocitrate lyase (ICL) and malate synthase (MLS) are two key enzymes that function exclusively in this pathway. Products of the glyoxylate cycle exit the peroxisome, enter gluconeogenesis, and are further converted to hexose and Suc to fuel the postgerminative development of seedlings (Penfield et al., 2006).Immunocytochemical studies of germinating seeds from pumpkin (Cucurbita pepo), watermelon (Citrullis vulgaris), and cucumber (Cucumis sativus) demonstrated that seed peroxisomes (glyoxysomes) are directly transformed into leaf peroxisomes during greening of the cotyledons without de novo biogenesis of leaf peroxisomes (Titus and Becker, 1985; Nishimura et al., 1986; Sautter, 1986). This conversion was illustrated by the import of photorespiratory enzymes and their concomitant presence with glyoxylate cycle enzymes within the same organelle. Furthermore, the increase in abundance of photorespiratory enzymes coincided with the marked decrease, and subsequently the absence, of glyoxylate cycle enzymes (ICL and/or MLS) at the culmination of this process (Titus and Becker, 1985; Nishimura et al., 1986; Sautter, 1986). It was suggested that the specific names for plant peroxisomal variants should be eliminated because protein composition between leaf peroxisomes and glyoxysomes may differ by only two proteins (i.e. ICL and MLS) out of the over 100 total proteins in the peroxisome (Pracharoenwattana and Smith, 2008). This prediction needed to be tested. In addition, mutants lacking core peroxisome biogenesis factors or major β-oxidation enzymes are nonviable, suggesting that peroxisomes are essential to embryogenesis and seed germination (Hu et al., 2012). However, how peroxisomes contribute to seed germination and seedling establishment is not completely understood. In the past, studies have been successfully undertaken to catalog the proteome of mitochondria and plastids isolated from different plant tissues, which uncovered unique facets of organelle metabolism in various tissues (van Wijk and Baginsky, 2011; Havelund et al., 2013; Lee et al., 2013). As such, it was necessary to establish a protein atlas for peroxisomes in dark-grown seedlings.Proteomic analyses of leaf peroxisomes and peroxisomes from suspension-cultured, leaf-derived cells followed by protein subcellular localization studies confirmed a total of over 30 new peroxisomal proteins, uncovering additional metabolic functions for leaf peroxisomes (Fukao et al., 2002; Reumann et al., 2007, 2009; Eubel et al., 2008; Babujee et al., 2010; Kataya and Reumann, 2010; Quan et al., 2010). For Arabidopsis (Arabidopsis thaliana), around 100 peroxisomal proteins were shown to be present in leaves or leaf-derived cells. Compared with the over 80 bona fide peroxisomal proteins detected by leaf peroxisomal proteomics (Reumann et al., 2007, 2009), the number of proteins identified from peroxisomal proteomic studies on etiolated seedlings was significantly smaller, with less than 10 known peroxisomal proteins from Arabidopsis (Fukao et al., 2003) and approximately 31 from soybean (Glycine max; Arai et al., 2008a, 2008b). Thus, a more in-depth analysis of the proteome of peroxisomes from these tissues was highly needed.Here, we performed proteomic analysis of peroxisomes isolated from etiolated Arabidopsis seedlings and detected peroxisomal proteins that encompass most of the known plant peroxisomal metabolic pathways. Fluorescence microscopy verified the peroxisomal localization of a number of proteins newly identified in this study or detected from previous proteomics that had not been verified by independent means. Reverse genetic analysis demonstrated the role for a Cys protease in germination, β-oxidation, and growth.  相似文献   
67.
Highlights? The GPCR Gpr161 localizes to primary cilia in a Tulp3/IFT-A-dependent manner ? Complete loss of Gpr161 increases Shh signaling in the mouse neural tube ? Gpr161 couples protein kinase A activation to Shh signaling via cAMP signaling ? Shh signaling internalizes Gpr161 from the cilia, preventing its activity  相似文献   
68.
CaSR and TGFβ are robust promoters of differentiation in the colonic epithelium. Loss of cellular responses to TGFβ or loss of CaSR expression is tightly linked to malignant progression. Human colonic epithelial CBS cells, originally developed from a differentiated human colon tumor, retain CaSR expression and function, TGFβ responsiveness and TGFβ receptor expression. Thus, these cells offer a unique opportunity in determining the functional linkage (if any) between CaSR and TGFβ. Knocking down CaSR expression abrogated TGFβ-mediated cellular responses and attenuated the expression of TGFβ receptors. Ca2+ or vitamin D treatment induced CaSR expression with a concurrent up-regulation of TGFβ receptor expression. Ca2+ or vitamin D, however, did not induce CaSR in CaSR knocked down cells and without CaSR; there was no up-regulation of TGFβ receptor. It is concluded that TGFβ receptor expression and TGFβ mediated responses requires CaSR expression and function.  相似文献   
69.
Tyagi NK  Fenton WA  Deniz AA  Horwich AL 《FEBS letters》2011,585(12):1969-1972
Under "permissive" conditions at 25°C, the chaperonin substrate protein DM-MBP refolds 5-10 times more rapidly in the GroEL/GroES folding chamber than in free solution. This has been suggested to indicate that the chaperonin accelerates polypeptide folding by entropic effects of close confinement. Here, using native-purified DM-MBP, we show that the different rates of refolding are due to reversible aggregation of DM-MBP while folding free in solution, slowing its kinetics of renaturation: the protein exhibited concentration-dependent refolding in solution, with aggregation directly observed by dynamic light scattering. When refolded in chloride-free buffer, however, dynamic light scattering was eliminated, refolding became concentration-independent, and the rate of refolding became the same as that in GroEL/GroES. The GroEL/GroES chamber thus appears to function passively toward DM-MBP.  相似文献   
70.
HIV-1 is neutralized by a class of antibodies that preferentially recognize a site formed on the assembled viral spike. Such quaternary structure-specific antibodies have diverse neutralization breadths, with antibodies PG16 and PG9 able to neutralize 70 to 80% of circulating HIV-1 isolates while antibody 2909 is specific for strain SF162. We show that alteration between a rare lysine and a common N-linked glycan at position 160 of HIV-1 gp120 is primarily responsible for toggling between 2909 and PG16/PG9 neutralization sensitivity. Quaternary structure-specific antibodies appear to target antigenic variants of the same epitope, with neutralization breadth determined by the prevalence of recognized variants among circulating isolates.  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