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   303篇
  免费   35篇
  338篇
  2023年   1篇
  2022年   1篇
  2021年   6篇
  2020年   8篇
  2019年   7篇
  2018年   6篇
  2017年   4篇
  2016年   10篇
  2015年   13篇
  2014年   13篇
  2013年   19篇
  2012年   20篇
  2011年   27篇
  2010年   16篇
  2009年   20篇
  2008年   16篇
  2007年   15篇
  2006年   19篇
  2005年   19篇
  2004年   20篇
  2003年   21篇
  2002年   19篇
  2001年   6篇
  2000年   2篇
  1999年   5篇
  1998年   2篇
  1997年   3篇
  1995年   5篇
  1994年   3篇
  1993年   2篇
  1991年   1篇
  1988年   1篇
  1984年   1篇
  1982年   1篇
  1981年   3篇
  1978年   1篇
  1976年   1篇
  1968年   1篇
排序方式: 共有338条查询结果,搜索用时 13 毫秒
61.
We present a comprehensive characterization of the nucleoside N-ribohydrolase (NRH) family in two model plants, Physcomitrella patens (PpNRH) and maize (Zea mays; ZmNRH), using in vitro and in planta approaches. We identified two NRH subclasses in the plant kingdom; one preferentially targets the purine ribosides inosine and xanthosine, while the other is more active toward uridine and xanthosine. Both subclasses can hydrolyze plant hormones such as cytokinin ribosides. We also solved the crystal structures of two purine NRHs, PpNRH1 and ZmNRH3. Structural analyses, site-directed mutagenesis experiments, and phylogenetic studies were conducted to identify the residues responsible for the observed differences in substrate specificity between the NRH isoforms. The presence of a tyrosine at position 249 (PpNRH1 numbering) confers high hydrolase activity for purine ribosides, while an aspartate residue in this position confers high activity for uridine. Bud formation is delayed by knocking out single NRH genes in P. patens, and under conditions of nitrogen shortage, PpNRH1-deficient plants cannot salvage adenosine-bound nitrogen. All PpNRH knockout plants display elevated levels of certain purine and pyrimidine ribosides and cytokinins that reflect the substrate preferences of the knocked out enzymes. NRH enzymes thus have functions in cytokinin conversion and activation as well as in purine and pyrimidine metabolism.Nucleoside hydrolases or nucleoside N-ribohydrolases (NRHs; EC 3.2.2.-) are glycosidases that catalyze the cleavage of the N-glycosidic bond in nucleosides to enable the recycling of the nucleobases and Rib (Fig. 1A). The process by which nucleosides and nucleobases are recycled is also known as salvaging and is a way of conserving energy, which would otherwise be needed for the de novo synthesis of purine- and pyrimidine-containing compounds. During the salvage, bases and nucleosides can be converted into nucleoside monophosphates by the action of phosphoribosyltransferases and nucleoside kinases, respectively, and further phosphorylated into nucleoside diphosphates and triphosphates (Moffatt et al., 2002; Zrenner et al., 2006; Fig. 1B). Uridine kinase and uracil phosphoribosyl transferase are key enzymes in the pyrimidine-salvaging pathway in plants (Mainguet et al., 2009; Chen and Thelen, 2011). Adenine phosphoribosyltransferase and adenosine kinase (ADK) are important in purine salvaging (Moffatt and Somerville, 1988; Moffatt et al., 2002), and their mutants cause reductions in fertility or sterility, changes in transmethylation, and the formation of abnormal cell walls. In addition, both enzymes were also reported to play roles in cytokinin metabolism (Moffatt et al., 1991, 2000; von Schwartzenberg et al., 1998; Schoor et al., 2011). Cytokinins (N6-substituted adenine derivatives) are plant hormones that regulate cell division and numerous developmental events (Mok and Mok, 2001; Sakakibara, 2006). Cytokinin ribosides are considered to be transport forms and have little or no activity.Open in a separate windowFigure 1.A, Scheme of the reactions catalyzed by plant NRHs when using purine (inosine), pyrimidine (uridine), and cytokinin (iPR) ribosides as the substrates. B, Simplified schematic overview of cytokinin, purine, and pyrimidine metabolism in plants. The diagram is adapted from the work of Stasolla et al. (2003) and Zrenner et al. (2006) with modifications. The metabolic components shown are as follows: 1, cytokinin nucleotide phosphoribohydrolase; 2, adenine phosphoribosyltransferase; 3, adenosine kinase; 4, 5′-nucleotidase; 5, adenosine phosphorylase; 6, purine/pyrimidine nucleoside ribohydrolase; 7, cytokinin oxidase/dehydrogenase; 8, AMP deaminase; 9, hypoxanthine phosphoribosyltransferase; 10, inosine kinase; 11, inosine-guanosine phosphorylase; 12, IMP dehydrogenase; 13, xanthine dehydrogenase; 14, 5′-nucleotidase; 15, GMP synthase; 16, hypoxanthine-guanine phosphoribosyltransferase; 17, guanosine deaminase; 18, guanine deaminase; 19, guanosine kinase; 20, uracil phosphoribosyltransferase; 21, uridine cytidine kinase; 22, pyrimidine 5′-nucleotidase; 23, cytidine deaminase; 24, adenosine/adenine deaminase. CK, Cytokinin; CKR, cytokinin riboside; CKRMP, cytokinin riboside monophosphate.NRHs are metalloproteins first identified and characterized in parasitic protozoa such as Trypanosoma, Crithidia, and Leishmania species that rely on the import and salvage of nucleotide derivatives. They have since been characterized in other organisms such as bacteria, yeast, and insects (Versées and Steyaert, 2003) but never in mammals (Parkin et al., 1991). They have been divided into four classes based on their substrate specificity: nonspecific NRHs, which hydrolyze inosine and uridine (IU-NRHs; Parkin et al., 1991; Shi et al., 1999); purine-specific inosine/adenosine/guanosine NRHs (Parkin, 1996); the 6-oxopurine-specific guanosine/inosine NRHs (Estupiñán and Schramm, 1994); and the pyrimidine nucleoside-specific cytidine/uridine NRHs (CU-NRHs; Giabbai and Degano, 2004). All NRHs exhibit a stringent specificity for the Rib moiety and differ in their preferences regarding the nature of the nucleobase. Crystal structures are available for empty NRH or in complex with inhibitors from Crithidia fasciculata (CfNRH; Degano et al., 1998), Leishmania major (LmNRH; Shi et al., 1999), and Trypanosoma vivax (TvNRH; Versées et al., 2001, 2002). The structures of two CU-NRHs from Escherichia coli, namely YeiK (Iovane et al., 2008) and YbeK (rihA; Muzzolini et al., 2006; Garau et al., 2010), are also available. NRHs are believed to catalyze N-glycosidic bond cleavage by a direct displacement mechanism. An Asp from a conserved motif acts as a general base and abstracts a proton from a catalytic water molecule, which then attacks the C1′ atom of the Rib moiety of the nucleoside. Kinetic isotope-effect studies on CfNRH (Horenstein et al., 1991) showed that the substrate’s hydrolysis proceeds via an oxocarbenium ion-like transition state and is preceded by protonation at the N7 atom of the purine ring, which lowers the electron density on the purine ring and destabilizes the N-glycosidic bond. A conserved active-site His is a likely candidate for this role in IU-NRHs and CU-NRHs. In the transition state, the C1′-N9 glycosidic bond is almost 2 Å long, with the C1′ atom being sp2 hybridized while the C3′ atom adopts an exo-conformation, and the whole ribosyl moiety carries a substantial positive charge (Horenstein et al., 1991).Several NRH enzymes have been identified in plants, including a uridine-specific NRH from mung bean (Phaseolus radiatus; Achar and Vaidyanathan, 1967), an inosine-specific NRH (EC 3.2.2.2) and a guanosine-inosine-specific NRH, both from yellow lupine (Lupinus luteus; Guranowski, 1982; Szuwart et al., 2006), and an adenosine-specific NRH (EC 3.2.2.7) from coffee (Coffea arabica), barley (Hordeum vulgare), and wheat (Triticum aestivum; Guranowski and Schneider, 1977; Chen and Kristopeit, 1981; Campos et al., 2005). However, their amino acid sequences have not been reported so far. A detailed study of the NRH gene family from Arabidopsis (Arabidopsis thaliana) has recently been reported (Jung et al., 2009, 2011). The AtNRH1 enzyme exhibits highest hydrolase activity toward uridine and xanthosine. It can also hydrolyze the cytokinin riboside N6-(2-isopentenyl)adenosine (iPR), which suggests that it may also play a role in cytokinin homeostasis. However, Riegler et al. (2011) analyzed the phenotypes of homozygous nrh1 and nrh2 single mutants along with the homozygous double mutants and concluded that AtNRHs are probably unimportant in cytokinin metabolism.Here, we identify and characterize plant IU-NRHs from two different model organisms, Physcomitrella patens and maize (Zea mays), combining structural, enzymatic, and in planta functional approaches. The moss P. patens was chosen to represent the bryophytes, which can be regarded as being evolutionarily basal terrestrial plants, and is suitable for use in developmental and metabolic studies (Cove et al., 2006; von Schwartzenberg, 2009), while maize is an important model system for cereal crops. We report the crystal structures of NRH enzymes from the two plant species, PpNRH1 and ZmNRH3. Based on these structures, we performed site-directed mutagenesis experiments and kinetic analyses of point mutants of PpNRH1 in order to identify key residues involved in nucleobase interactions and catalysis. To analyze the physiological role of the PpNRHs, single knockout mutants were generated. NRH deficiency caused significant changes in the levels of purine, pyrimidine, and cytokinin metabolites relative to those seen in the wild type, illustrating the importance of these enzymes in nucleoside and cytokinin metabolism.  相似文献   
62.
Fgf8 signalling is known to play an important role during patterning of the first pharyngeal arch, setting up the oral region of the head and then defining the rostral and proximal domains of the arch. The mechanisms that regulate the restricted expression of Fgf8 in the ectoderm of the developing first arch, however, are not well understood. It has become apparent that pharyngeal endoderm plays an important role in regulating craniofacial morphogenesis. Endoderm ablation in the developing chick embryo results in a loss of Fgf8 expression in presumptive first pharyngeal arch ectoderm. Shh is locally expressed in pharyngeal endoderm, adjacent to the Fgf8-expressing ectoderm, and is thus a candidate signal regulating ectodermal Fgf8 expression. We show that in cultured explants of presumptive first pharyngeal arch, loss of Shh signalling results in loss of Fgf8 expression, both at early stages before formation of the first arch, and during arch formation. Moreover, following removal of the endoderm, Shh protein can replace this tissue and restore Fgf8 expression. Overexpression of Shh in the non-oral ectoderm leads to an expansion of Fgf8, affecting the rostral-caudal axis of the developing first arch, and resulting in the formation of ectopic cartilage. Shh from the pharyngeal endoderm thus regulates Fgf8 in the ectoderm and the role of the endoderm in pharyngeal arch patterning may thus be indirectly mediated by the ectoderm.  相似文献   
63.
The paleoecological interpretation of fossil foraminiferal assemblages depends on an understanding of the ecological processes operating at the present. This study investigates both the quality of organic matter (OM) by elemental analysis as well as the sediment grain size and clay mineralogy to understand their relative influence on distribution and abundance of benthic foraminifera. This study is carried out on 15 samples regularly spaced from the mudflat to the tidal marsh. The results indicate that grain size is the most limiting parameter. Living (stained) benthic foraminiferal density and species richness are both very low within coarser sediments. OM is the second limiting factor. The density of foraminifera is the lowest and the species richness is the highest with the lowest organic carbon (Corg) contents and C/N < 12. Conversely, when the Corg is very high and C/N > 12, the density is high and the species richness medium. A high smectite proportion within the clay-size fraction seems to favor the development of Miliammina fusca. Trochammina inflata and Jadammina macrescens are both favored by an increase of organic carbon proportion but Trochammina inflata preferentially feeds on algal-derived OM when compared with Jadammina macrescens.  相似文献   
64.
65.
66.
During pregnancy the decline in blood [glucose] does not result from the increased distribution space of glucose. The absolute rate of glucose turnover increases in late pregnancy in parallel with the rise in the mass of the conceptus. Nevertheless, glucose turnover per kg body wt. is not increased in late pregnancy, since the lower blood [glucose] decreases glucose utilization by maternal tissues.  相似文献   
67.
68.

Background  

Most aphid species complete their life cycle on the same set of host-plant species, but some (heteroecious species) alternate between different hosts, migrating from primary (woody) to secondary (herbaceous) host plants. The evolutionary processes behind the evolution of this complex life cycle have often been debated. One widely accepted scenario is that heteroecy evolved from monoecy on woody host plants. Several shifts towards monoecy on herbaceous plants have subsequently occurred and resulted in the radiation of aphids. Host alternation would have persisted in some cases due to developmental constraints preventing aphids from shifting their entire life cycle to herbaceous hosts (which are thought to be more favourable). According to this scenario, if aphids lose their primary host during evolution they should not regain it. The genus Brachycaudus includes species with all the types of life cycle (monoecy on woody plants, heteroecy, monoecy on herbs). We used this genus to test hypotheses concerning the evolution of life cycles in aphids.  相似文献   
69.
The initiation of eukaryotic DNA replication is a highly regulated process conserved from yeast to human. The past decade has seen significant advances in understanding how the CMG (Cdc45‐MCM‐GINS) replicative helicase is loaded onto DNA. However, very little was known on how this complex is removed from chromatin at the end of S phase. Two papers in a recent issue of Science 1 2 show that in yeast and in Xenopus, the CMG complex is unloaded at replication termination sites by an active mechanism involving the polyubiquitylation of Mcm7.  相似文献   
70.
This study aimed to understand the response of photosynthesis and growth to e-CO2 conditions (800 vs. 400 μmol mol−1) of rice genotypes differing in source–sink relationships. A proxy trait called local C source–sink ratio was defined as the ratio of flag leaf area to the number of spikelets on the corresponding panicle, and five genotypes differing in this ratio were grown in a controlled greenhouse. Differential CO2 resources were applied either during the 2 weeks following heading (EXP1) or during the whole growth cycle (EXP2). Under e-CO2, low source–sink ratio cultivars (LSS) had greater gains in photosynthesis, and they accumulated less nonstructural carbohydrate in the flag leaf than high source–sink ratio cultivars (HSS). In EXP2, grain yield and biomass gain was also greater in LSS probably caused by their strong sink. Photosynthetic capacity response to e-CO2 was negatively correlated across genotypes with local C source–sink ratio, a trait highly conserved across environments. HSS were sink-limited under e-CO2, probably associated with low triose phosphate utilization (TPU) capacity. We suggest that the local C source–sink ratio is a potential target for selecting more CO2-responsive cultivars, pending validation for a broader genotypic spectrum and for field conditions.  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