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   1448篇
  免费   69篇
  1517篇
  2023年   6篇
  2022年   7篇
  2021年   11篇
  2020年   13篇
  2019年   10篇
  2018年   22篇
  2017年   13篇
  2016年   22篇
  2015年   46篇
  2014年   47篇
  2013年   117篇
  2012年   74篇
  2011年   83篇
  2010年   65篇
  2009年   51篇
  2008年   100篇
  2007年   71篇
  2006年   74篇
  2005年   77篇
  2004年   127篇
  2003年   102篇
  2002年   70篇
  2001年   21篇
  2000年   19篇
  1999年   18篇
  1998年   20篇
  1997年   11篇
  1996年   15篇
  1995年   8篇
  1994年   12篇
  1993年   14篇
  1992年   15篇
  1991年   18篇
  1990年   18篇
  1989年   9篇
  1988年   11篇
  1987年   4篇
  1986年   7篇
  1985年   13篇
  1984年   14篇
  1983年   6篇
  1982年   4篇
  1980年   6篇
  1979年   4篇
  1978年   6篇
  1977年   6篇
  1976年   8篇
  1975年   5篇
  1973年   4篇
  1969年   3篇
排序方式: 共有1517条查询结果,搜索用时 15 毫秒
51.
Ferredoxin (Fd) is the major iron-containing protein in photosynthetic organisms and is central to reductive metabolism in the chloroplast. The Chlamydomonas reinhardtii genome encodes six plant type [Fe2S2] ferredoxins, products of PETF, FDX2–FDX6. We performed the functional analysis of these ferredoxins by localizing Fd, Fdx2, Fdx3, and Fdx6 to the chloroplast by using isoform-specific antibodies and monitoring the pattern of gene expression by iron and copper nutrition, nitrogen source, and hydrogen peroxide stress. In addition, we also measured the midpoint redox potentials of Fd and Fdx2 and determined the kinetic parameters of their reactions with several ferredoxin-interacting proteins, namely nitrite reductase, Fd:NADP+ oxidoreductase, and Fd:thioredoxin reductase. We found that each of the FDX genes is differently regulated in response to changes in nutrient supply. Moreover, we show that Fdx2 (Em = −321 mV), whose expression is regulated by nitrate, is a more efficient electron donor to nitrite reductase relative to Fd. Overall, the results suggest that each ferredoxin isoform has substrate specificity and that the presence of multiple ferredoxin isoforms allows for the allocation of reducing power to specific metabolic pathways in the chloroplast under various growth conditions.Ferredoxins are small (∼11,000-kDa), soluble, iron-sulfur cluster-containing proteins with strongly negative redox potentials (−350 to −450 mV) that function as electron donors at reductive steps in various metabolic pathways (13). In photosynthetic organisms, the well studied ferredoxin (Fd4; the product of the PETF gene) is the most abundant iron-containing protein in the chloroplast and is central to the distribution of photosynthetically derived reductive power (4).The most well known Fd-dependent reaction is the transfer of electrons from photosystem I (PSI) to NADPH, catalyzed by Fd:NADP+ oxidoreductase (FNR). The NADPH produced by this reaction donates electrons to the only reductant-requiring step in the Calvin cycle and other steps in anabolic pathways that require NADPH as reductant. In addition, reduced Fd directly donates electrons to other metabolic pathways by interacting with various enzymes in the chloroplast. This includes Fd:thioredoxin reductase (FTR), which converts a light-driven electron signal into a thiol signal that is transmitted to thioredoxins (TRXs) present in the plastid as different types (or different isoforms). Once reduced, TRXs interact with specific disulfide bonds on target enzymes, modulating their activities (5). Other Fd targets include hydrogenase, which is responsible for hydrogen production in anaerobic conditions in green algae; glutamine-oxoglutarate amidotransferase in amino acid synthesis; nitrite and sulfite reductases in nitrate and sulfate assimilation, respectively; stearoyl-ACP Δ9-desaturase in fatty acid desaturation; and phycocyanobilin:Fd oxidoreductase in synthesis of phytochromobilin (6). Fd also functions in non-photosynthetic cells. Here, FNR catalyzes the reduction of Fd by NADPH produced in the oxidative pentose phosphate pathway, enabling Fd-dependent metabolism to occur in the dark (7, 8).The single-celled green alga, Chlamydomonas reinhardtii is an excellent reference organism for studying both metabolic adaptation to nutrient stress and photosynthesis (913). The Chlamydomonas genome encodes six highly related plant type ferredoxin genes (9). Until recently, only the major photosynthetic ferredoxin, Fd (encoded by PETF), which mediates electron transfer between PSI and FNR, had been characterized in detail (14).Many land plants are known to have multiple ferredoxins. Typically, they are differently localized on the basis of their function. Photosynthetic ferredoxins reduce NADP+ at a faster rate and are localized to the leaves, whereas non-photosynthetic ferredoxins are more efficiently reduced by NADPH and are localized to the roots. Arabidopsis thaliana has a total of six ferredoxin isoforms (15). Of these, two are photosynthetic and localized in the leaves. The most abundant, AtFd2, is involved in linear electron flow, and the less abundant (5% of the ferredoxin pool), AtFd1, has been implicated in cyclic electron flow (16). There is one non-photosynthetic ferredoxin located in the roots, AtFd3, which is nitrate-inducible. This protein has higher electron transfer activity with sulfite reductase in in vitro assays compared with other Arabidopsis ferredoxin isoforms, suggesting in vivo function of AtFd3 in nitrate and sulfate assimilation (15, 17). In addition, there is one evolutionarily distant ferredoxin, AtFd4, of unknown function with a more positive redox potential present in the leaves and two other proteins which are “ferredoxin-like” and uncharacterized (15). Zea mays has four ferredoxin isoforms, two photosynthetic and two non-photosynthetic (18). One of the non-photosynthetic isoforms is specifically induced by nitrite, suggestive of a role in nitrate metabolism (19). A cyanobacterium, Anabaena 7120, has two ferredoxins, vegetative and heterocyst type (by analogy to leaf and root types, respectively). The heterocyst type is present only in cells that have differentiated into nitrogen-fixing cells, indicating that this form may serve to transfer electrons to nitrogenase (20).We hypothesize that the presence of as many as six ferredoxin isoforms in a single-celled organism like C. reinhardtii allows for the differential regulation of each isoform and therefore the prioritization of reducing power toward certain metabolic pathways under changing environmental conditions. To test this hypothesis, expression of the genes (PETF and FDX2–FDX6) encoding the six ferredoxin isoforms in Chlamydomonas reinhardtii was monitored under various conditions in which well characterized ferredoxin-dependent enzymes are known to be expressed. In addition, we also analyzed the interaction of Fd and Fdx2 with several ferredoxin-interacting proteins, such as NiR, FNR, and FTR, and determined the kinetic parameters of the corresponding reactions.We found that each of the FDX genes is indeed differently regulated in response to changes in nutrient supply. In the case of FDX2 whose product is most similar to classical Fd, we suggest that it has specificity for nitrite reductase based on its pattern of expression and activity with nitrite reductase.  相似文献   
52.
Epigenetic changes are thought to lead to alterations in the property of cells, such as differentiation potential. Neural precursor cells (NPCs) differentiate only into neurons in the midgestational brain, yet they become able to generate astrocytes in the late stage of development. This differentiation-potential switch could be explained by epigenetic changes, since the promoters of astrocyte-specific marker genes, glial fibrillary acidic protein (Gfap) and S100beta, have been shown to become demethylated in late-stage NPCs prior to the onset of astrocyte differentiation; however, whether demethylation occurs generally in other astrocyctic genes remains unknown. Here we analyzed DNA methylation changes in mouse NPCs between the mid-(E11.5) and late (E14.5) stage of development by a genome-wide DNA methylation profiling method using microarrays and found that many astrocytic genes are demethylated in late-stage NPCs, enabling the cell to become competent to express these genes. Although these genes are already demethylated in late-stage NPCs, they are not expressed until cells differentiate into astrocytes. Thus, late-stage NPCs have epigenetic potential which can be realized in their expression after astrocyte differentiation.  相似文献   
53.

Background  

The mosquito Anopheles irenicus, a member of the Anopheles punctulatus group, is geographically restricted to Guadalcanal in the Solomon Islands. It shows remarkable morphological similarities to one of its sibling species, An. farauti sensu stricto (An. farauti s.s.), but is dissimilar in host and habitat preferences. To infer the genetic variations between these two species, we have analyzed mitochondrial cytochrome oxidase subunit II (COII) and nuclear ribosomal internal transcribed spacer 2 (ITS2) sequences from Guadalcanal and from one of its nearest neighbours, Malaita, in the Solomon Islands.  相似文献   
54.
Endothelin (ET)-1 is a potent inducer of peptic ulcers. The roles of ET-1 in ulcer healing, however, have remained unclear, and these were investigated in mice. Gastric ulcers were induced in mice by serosal application of acetic acid. Three days later, mice were given a neutralizing ET-1 antibody or nonimmunized serum. The ulcer size, amount of fibrosis and myofibroblasts, and localization of ET-1 and ET(A/B) receptors were analyzed. To elucidate the mechanisms underlying the effects of ET-1, we examined the proliferation, migration, and release of growth and angiogenic factors in gastric myofibroblasts with or without ET-1. The expression of prepro-ET-1 (an ET-1 precursor) and ET-converting enzyme-1 was examined in gastric myofibroblasts using RT-PCR. Immunoneutralization of ET-1 delayed gastric ulcer healing. The areas of fibrosis and myofibroblasts were smaller in the anti-ET-1 antibody group than in the control. ET-1 was expressed in the gastric epithelium, myofibroblasts, and other cell types. ET(A) receptors, but not ET(B) receptors, were present in myofibroblasts. ET-1 increased proliferation and migration of gastric myofibroblasts. ET-1 stimulated the release of hepatocyte growth factor, VEGF, PGE(2), and IL-6 from gastric myofibroblasts. mRNA for prepro-ET-1 and ET-converting enzyme-1 was also expressed. ET-1 promotes the accumulation of gastric myofibroblasts and collagen fibrils at gastric ulcers. ET-1 also stimulates migration and proliferation of gastric myofibroblasts and enhances the release of growth factors, angiogenic factors, and PGE(2). Thus ET-1 has important roles not only in ulcer formation but also in ulcer healing via mobilizing myofibroblasts and inducing production of stroma-derived factors.  相似文献   
55.
In the continuing study directed toward the development of peroxisome proliferator-activated receptor gamma (hPPARγ) agonist, we attempted to improve the water solubility of our previously developed hPPARγ-selective agonist 3, which is insufficiently soluble for practical use, by employing two strategies: introducing substituents to reduce its molecular planarity and decreasing its hydrophobicity via replacement of the adamantyl group with a heteroaromatic ring. The first approach proved ineffective, but the second was productive. Here, we report the design and synthesis of a series of α-benzyl phenylpropanoic acid-type hPPARγ partial agonists with improved aqueous solubility. Among them, we selected (R)-7j, which activates hPPARγ to the extent of about 65% of the maximum observed with a full agonist, for further evaluation. The ligand-binding mode and the reason for the partial-agonistic activity are discussed based on X-ray-determined structure of the complex of hPPARγ ligand-binding domain (LBD) and (R)-7j with previously reported ligand-LDB structures. Preliminal apoptotic effect of (R)-7j against human scirrhous gastric cancer cell line OCUM-2MD3 is also described.  相似文献   
56.
The content of total cellular lipid of Candida tropicalis grown on a mixture of n-alkanes (C10–C18) was about 20% of the dry cell weight at the exponential growth phase and 14% at the early stationary phase. Phospholipid corresponded to approximately 70 % of the total lipid independent of the growth phases. The composition of cellular lipid classes did not change significantly during the growth. On the other hand, a drastic time-course change in fatty acid composition was observed. The proportion of odd-chain fatty acids, one of the most specific cellular components of the yeast grown on the n-alkane mixture, increased in both phospholipid and triglyceride along with the yeast growth. In the meantime, the proportion of polyunsaturated fatty acids varied markedly during the course of cultivation, showing a peak at the early growth phase. The high content of polyunsaturated fatty acids at the early stages of growth correlated to the contents of these acids in phospholipid rather than in triglyceride.  相似文献   
57.
58.
Galectin-9 ameliorates various murine autoimmune disease models by regulating T cells and macrophages, although it is not known what role it may have in B cells. The present experiment shows that galectin-9 ameliorates a variety of clinical symptoms, such as proteinuria, arthritis, and hematocrit in MRL/lpr lupus-prone mice. As previously reported, galectin-9 reduces the frequency of Th1, Th17, and activated CD8+ T cells. Although anti-dsDNA antibody was increased in MRL/lpr lupus-prone mice, galectin-9 suppressed anti-dsDNA antibody production, at least partly, by decreasing the number of plasma cells. Galectin-9 seemed to decrease the number of plasma cells by inducing plasma cell apoptosis, and not by suppressing BAFF production. Although about 20% of CD19−/low CD138+ plasma cells expressed Tim-3 in MRL/lpr lupus-prone mice, Tim-3 may not be directly involved in the galectin-9-induced apoptosis, because anti-Tim-3 blocking antibody did not block galectin-9-induced apoptosis. This is the first report of plasma cell apoptosis being induced by galectin-9. Collectively, it is likely that galectin-9 attenuates the clinical severity of MRL lupus-prone mice by regulating T cell function and inducing plasma cell apoptosis.  相似文献   
59.
The present study examined the cellular functions of low-molecular-weight protein tyrosine phosphatase (LMW-PTP), which consists of two active isoforms IF-1 and IF-2, in vascular smooth muscle cells (VSMCs) and endothelial cells (ECs), focusing on cell growth and migration. We transduced recombinant IF-1 and IF-2, and ribozyme targeting both isoforms using an adenovirus vector in these cells. We detected the expression of IF-1 and IF-2 in both types of cells. IF-1 as well as IF-2 inhibited PDGF-induced DNA synthesis and migration in VSMCs. In contrast, both isoforms enhanced lysophosphatidic acid-stimulated cell migration without change in DNA synthesis in ECs. Whereas there is a report indicating that reactive oxygen species-dependent inactivation of LMW-PTP regulates actin cytoskeleton reorganization during cell spreading and migration, the isoforms conversely suppressed the PDGF-induced H2O2 generation with subsequent decrease in the p38 activity in VSMCs. Catalytically inactive LMW-PTP exerted the opposite and similar effects to the wild type in ECs and in VSMCs, respectively, suggesting that substrates for the phosphatase differ between these cells. Moreover, high concentrations of glucose suppressed the expression of LMW-PTP in both cells. These data suggest that LMW-PTP negatively regulates the pathogenesis of atherosclerosis and that glucose-dependent suppression of LMW-PTP expression may promote the development of atherosclerosis in diabetics.  相似文献   
60.
In the contemporary era, when life habits are largely determined by social needs and individual preferences, sleep is nevertheless affected by seasonal environmental changes. Japan has large seasonal and geographical alterations of photoperiod and climate. Japan does not adopt the daylight saving time (DST) system, making it a suitable country for the study of seasonal variations in natural human sleep. The aim of this study was to analyze the seasonal changes in the sleep properties (timing and quality) and identify their relationship with environmental changes. Here, we report an analysis of objective sleep data of 691 161 nights collected from 1856 Japanese participants (age 20–79 years, male 91%, female 9%) for 3 years using contactless biomotion sensors. Sleep onset time did not show clear seasonal variation, but sleep offset time showed a seasonal change with a single latest peak in winter. Seasonal variation was larger during weekends than during weekdays. Sleep offset time well correlated with sunrise time but was different in spring and autumn even when the sunrise time was same, suggesting the role of temperature difference. Sleep quality, estimated by wake time after sleep onset and sleep efficiency, showed seasonal changes with the lowest trough around mid-summer. In conclusion, despite profound social influences, the timing and quality of sleep showed seasonal fluctuation indicating that they were influenced by climate factors even in the developed country.  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