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   2113篇
  免费   200篇
  2022年   17篇
  2021年   33篇
  2020年   18篇
  2019年   29篇
  2018年   28篇
  2017年   27篇
  2016年   49篇
  2015年   83篇
  2014年   89篇
  2013年   133篇
  2012年   143篇
  2011年   122篇
  2010年   79篇
  2009年   73篇
  2008年   103篇
  2007年   111篇
  2006年   92篇
  2005年   98篇
  2004年   71篇
  2003年   72篇
  2002年   74篇
  2001年   42篇
  2000年   48篇
  1999年   33篇
  1998年   28篇
  1997年   14篇
  1996年   20篇
  1995年   14篇
  1993年   19篇
  1992年   28篇
  1991年   23篇
  1990年   27篇
  1989年   32篇
  1988年   27篇
  1987年   28篇
  1986年   22篇
  1985年   24篇
  1984年   25篇
  1983年   18篇
  1982年   12篇
  1981年   14篇
  1980年   11篇
  1979年   20篇
  1978年   13篇
  1977年   13篇
  1976年   14篇
  1975年   18篇
  1973年   14篇
  1971年   12篇
  1969年   10篇
排序方式: 共有2313条查询结果,搜索用时 15 毫秒
991.
Peyer's patch lymphocytes cocultured with Caco-2 cells trigger the phenotypic conversion of enterocytes into cells that express morphological and functional M-cell properties. We report a technical update for setting up this model, which will enable the study of M-cell biology, the identification by biochemical approaches of molecules involved in the interaction of microorganisms with M cells, and the development of vectors that would efficiently target the mucosal immune system.  相似文献   
992.
The contribution of N2 fixation to overall soybean N uptake has most commonly been quantified by N isotope‐based methods, which rely on isotopic differences in plant N between legumes and non‐fixing reference plants. The choice of non‐fixing reference plants is critical for the accuracy of isotope‐based methods, and mismatched reference plants remain a potential source of error. Accurate estimates of soybean N2 fixation also require information on N isotopic fractionation within soybean. On the basis of a previous observation of a close correlation between an expression of N fractionation within soybean and the proportion of plant N derived from atmosphere (%Ndfa) determined by 15N natural abundance, this field study aimed at assessing the relationship between various expressions describing intraplant 15N or N partitioning and %Ndfa during soybean development. Starting from a late vegetative stage until beginning senescence, the N content and N isotopic composition of shoots, roots and nodules of nodulated and non‐nodulated soybeans was determined at eight different developmental stages. Regression analysis showed that %Ndfa most closely correlated with the difference in the N isotopic composition of shoot N minus that of root including nodule N, and that this relationship was similar to that obtained in a previous multi‐site field study. We therefore consider this expression to hold promise as a means of quantifying %Ndfa independent of a reference plant, which would avoid some of the external sources of error introduced by the use of reference plants in determining %Ndfa.  相似文献   
993.
Lipid-linked oligosaccharides (LLOs) are the substrates of oligosaccharyltransferase (OST), the enzyme that catalyzes the en bloc transfer of the oligosaccharide onto the acceptor asparagine of nascent proteins during the process of N-glycosylation. To explore LLOs’ preferred location, orientation, structure, and dynamics in membrane bilayers of three different lipid types (dilauroylphosphatidylcholine, dimyristoylphosphatidylcholine, and dioleoylphosphatidylcholine), we have modeled and simulated both eukaryotic (Glc3-Man9-GlcNAc2-PP-Dolichol) and bacterial (Glc1-GalNAc5-Bac1-PP-Undecaprenol) LLOs, which are composed of an isoprenoid moiety and an oligosaccharide, linked by pyrophosphate. The simulations show no strong impact of different bilayer hydrophobic thicknesses on the overall orientation, structure, and dynamics of the isoprenoid moiety and the oligosaccharide. The pyrophosphate group stays in the bilayer head group region. The isoprenoid moiety shows high flexibility inside the bilayer hydrophobic core, suggesting its potential role as a tentacle to search for OST. The oligosaccharide conformation and dynamics are similar to those in solution, but there are preferred interactions between the oligosaccharide and the bilayer interface, which leads to LLO sugar orientations parallel to the bilayer surface. Molecular docking of the bacterial LLO to a bacterial OST suggests that such orientations can enhance binding of LLOs to OST.  相似文献   
994.
995.
The ability of retroviruses and transposons to insert their genetic material into host DNA makes them widely used tools in molecular biology, cancer research and gene therapy. However, these systems have biases that may strongly affect research outcomes. To address this issue, we generated very large datasets consisting of to unselected integrations in the mouse genome for the Sleeping Beauty (SB) and piggyBac (PB) transposons, and the Mouse Mammary Tumor Virus (MMTV). We analyzed (epi)genomic features to generate bias maps at both local and genome-wide scales. MMTV showed a remarkably uniform distribution of integrations across the genome. More distinct preferences were observed for the two transposons, with PB showing remarkable resemblance to bias profiles of the Murine Leukemia Virus. Furthermore, we present a model where target site selection is directed at multiple scales. At a large scale, target site selection is similar across systems, and defined by domain-oriented features, namely expression of proximal genes, proximity to CpG islands and to genic features, chromatin compaction and replication timing. Notable differences between the systems are mainly observed at smaller scales, and are directed by a diverse range of features. To study the effect of these biases on integration sites occupied under selective pressure, we turned to insertional mutagenesis (IM) screens. In IM screens, putative cancer genes are identified by finding frequently targeted genomic regions, or Common Integration Sites (CISs). Within three recently completed IM screens, we identified 7%–33% putative false positive CISs, which are likely not the result of the oncogenic selection process. Moreover, results indicate that PB, compared to SB, is more suited to tag oncogenes.  相似文献   
996.
Proteomic changes of two types of human endothelial cells (ECs) were determined and compared to morphological alterations occurring during the scaffold‐free in vitro formation of 3D structures resembling vascular intimas. The EA.hy926 cell line and human microvascular ECs (HMVECs) were cultured on a random positioning machine or static on ground (normal gravity) for 5 and 7 days, before their morphology was examined and their protein content was analysed by MS after free‐flow electrophoretic separation. A total of 1175 types of proteins were found in EA.hy926 cells and 846 in HMVEC forming 3D structures faster than the EA.hy926 cells. Five hundred and eighty‐four of these kinds of proteins were present in both types of cells. They included a number of metabolic enzymes, of structure‐related and stress proteins. Comparing proteins of EA.hy926 cells growing either adherently on ground or in 3D aggregates on the random positioning machine revealed that ribosomal proteins were enhanced, while tubes are formed and various components of 26S proteasomes remained prevalent in static normal gravity control cells only. The fast developing tube‐like 3D structures of HMVEC suggested a transient augmentation of ribosomal proteins during the 3D assembling of ECs.  相似文献   
997.
998.
The G protein-coupled receptor Gpr30 (Gper) was recently claimed to bind to estradiol and to activate cytoplasmic signal transduction pathways in response to estradiol. However, there are conflicting data regarding the role of Gpr30 as an estrogen receptor (ER): several laboratories were unable to demonstrate estradiol binding to GPR30 or estradiol-activated signal transduction in Gpr30-expressing cells. To clarify the potential role of Gpr30 as an ER, we generated Gpr30-deficient mice. Although Gpr30 was expressed in all reproductive organs, histopathological analysis did not reveal any abnormalities in these organs in Gpr30-deficient mice. Mutant male and female mice were as fertile as their wild-type littermates, indicating normal function of the hypothalamic-pituitary-gonadal axis. Moreover, we analyzed estrogenic responses in two major estradiol target organs, the uterus and the mammary gland. For that purpose, we examined different readout paradigms such as morphological measures, cellular proliferation, and target gene expression. Our data demonstrate that in vivo Gpr30 is dispensable for the mediation of estradiol effects in reproductive organs. These results are in clear contrast to the phenotype of mice lacking the classic ER alpha (Esr1) or aromatase (Cyp19a1). We conclude that the perception of Gpr30 (based on homology related to peptide receptors) as an ER might be premature and has to be reconsidered.  相似文献   
999.
DivIB(FtsQ), FtsL, and DivIC(FtsB) are enigmatic membrane proteins that are central to the process of bacterial cell division. DivIB(FtsQ) is dispensable in specific conditions in some species, and appears to be absent in other bacterial species. The presence of FtsL and DivIC(FtsB) appears to be conserved despite very low sequence conservation. The three proteins form a complex at the division site, FtsL and DivIC(FtsB) being associated through their extracellular coiled-coil region. We report here structural investigations by NMR, small-angle neutron and x-ray scattering, and interaction studies by surface plasmon resonance, of the complex of DivIB, FtsL, and DivIC from Streptococcus pneumoniae, using soluble truncated forms of the proteins. We found that one side of the “bean”-shaped central β-domain of DivIB interacts with the C-terminal regions of the dimer of FtsL and DivIC. This finding is corroborated by sequence comparisons across bacterial genomes. Indeed, DivIB is absent from species with shorter FtsL and DivIC proteins that have an extracellular domain consisting only of the coiled-coil segment without C-terminal conserved regions (Campylobacterales). We propose that the main role of the interaction of DivIB with FtsL and DivIC is to help the formation, or to stabilize, the coiled-coil of the latter proteins. The coiled-coil of FtsL and DivIC, itself or with transmembrane regions, could be free to interact with other partners.Cell division is one of the defining features of life. Understanding the division of bacteria is also required to find novel antibiotic strategies. Numerous studies, carried out mostly with the model organisms Escherichia coli and Bacillus subtilis have uncovered several components of the divisome, which can be defined as the ensemble of proteins localized at the division site and participating in the process. Comparison of genomes and deletion studies indicate that the core of the divisome comprises eight conserved, mostly essential proteins: FtsZ, FtsA, FtsK, FtsQ(DivIB), FtsL, FtsB(DivIC), FtsW, and FtsI. Fts nomenclature applies to Gram-negative organisms, whereas Div nomenclature applies to Gram-positive bacteria. These proteins are listed here in the conditional order of their recruitment to the division site of E. coli (14).Processes in which they participate have been attributed to several division proteins. FtsZ forms polymers with an annular distribution on the cytoplasmic side of the membrane and governs the recruitment of the other proteins. FtsA may mediate the interaction of FtsZ with the membrane. FtsK participates to the resolution of chromosome dimers, and possibly to the membrane fission. FtsI, and likely FtsW, participate to septal cell wall formation (14). In contrast, the roles of FtsQ(DivIB), FtsL, and FtsB(DivIC) have not been firmly linked to any particular process.FtsQ(DivIB), FtsL, and FtsB(DivIC) are positioned in the middle of the conditional order of recruitment in E. coli and B. subtilis. When the temporality of the recruitment was examined, FtsQ(DivIB) was found to belong to the late recruits, together with the proteins involved in cell wall assembly (5). In E. coli, the presence of FtsL and FtsB at the division site is mutually dependent, and their localization depends on that of FtsQ (6, 7). In B. subtilis, the presence of FtsL and DivIC at mid-cell depends on that of DivIB, at the temperature at which DivIB is essential, and reciprocally (8, 9). A complex comprising FtsQ, FtsL, and FtsB was isolated from E. coli by co-immunoprecipitation (10), and reconstituted in vitro with recombinant soluble forms of pneumococcal DivIB, FtsL, and DivIC (11). The interaction of the three proteins was also confirmed by yeast and bacterial triple hybrid (12, 13).The genes ftsL and ftsB(divIC) are essential in E. coli and B. subtilis (6, 1416) and presumably in Streptococcus pneumoniae (17). The essentiality of ftsQ(divIB) in laboratory conditions varies between species. The gene ftsQ is essential in E. coli (18), but divIB is essential only at high temperatures in B. subtilis (9, 19), or in a chemically defined medium in S. pneumoniae (17). Under these conditions, the essentiality of DivIB appears to be a consequence of the protection from proteolysis that it provides to FtsL (8, 17).FtsQ(DivIB), FtsL, and FtsB(DivIC) are bitopic membrane proteins with an N-terminal cytoplasmic region, a single transmembrane segment, and an extracytoplasmic region. The extracellular part is necessary and sufficient for the localization and function of FtsQ(DivIB), provided that it is anchored to the membrane (e.g. Refs. 20 and 21)), although the transmembrane segment also contributes to the localization (22, 23). The extracellular part is organized in three regions termed α, β, and γ. The crystal structure of a region consisting of the α- and β-domains was solved for FtsQ from E. coli and Yersinia enterocolitica (24). The α-domain, comprising about 70 amino acids proximal to the cytoplasmic membrane, corresponds to the POTRA (for polypeptide transport-associated) domain first identified by sequence analysis and proposed to function as a molecular chaperone (25). The α- and β-domains form the conserved region of the FtsQ(DivIB) protein. The γ-region constitutes a C-terminal tail. It is highly variable in length and sequence and predicted to be unfolded. The γ-region was not observed in the structures from E. coli and Y. entercolitica, thus confirming its flexible nature (24).The α-domain in the recombinant soluble form of the extracellular part of DivIB from Geobacillus stearothermophilus was digested by trypsin and therefore considered to be largely unfolded (26). The γ-region was also removed by trypsin digestion, together with a C-terminal fragment of the β-domain. The structure of the resulting shorter β-domain from G. stearothermophilus was solved by NMR (26) and lacks the two C-terminal β-strands.Localization epitopes have been identified in the transmembrane segment, the α-domain, and a region encompassing the C-terminal part of the β-domain and γ-tail of DivIB from B. subtilis (23). Likewise in E. coli, a region in the α-domain is required for localization of FtsQ, whereas the C-terminal region of the β-domain and the last α-helix are required for recruitment of FtsL and FtsB (24). In S. pneumoniae, the essentiality of DivIB in defined medium was found to reside in the C-terminal region of the β-domain (17).No experimental structure is known for FtsL or FtsB(DivIC). Both are small proteins comprising between 90 and 140 amino acids. The number of residues is sometimes larger, as in Mycobacterium tuberculosis (384 for FtsL and 228 for FtsB), due to N- and/or C-terminal extensions consisting of mostly charged and polar amino acids or proline-rich sequences. The major part of FtsL or FtsB(DivIC) is extracellular and contains a region proximal to the transmembrane segment, predicted to form a coiled-coil of about five heptads. Coiled-coil helices associate longitudinally to mediate protein association. It is possible that the coiled-coil helices are continuations of the transmembrane helices, although a proline (known to break helices) is present in some species between the two segments. Following the coiled-coil region is a 25–35-residue long C-terminal region in both FtsL and DivIC(FtsB). This region was recently shown in FtsB to be required for interaction with FtsQ in E. coli (27).We report here the results of structural studies in solution of a ternary complex consisting of the β- and γ-segments of DivIB, and a constrained dimer of the extracellular parts of FtsL and DivIC from S. pneumoniae. Despite the coiled-coil predictions, the recombinant extracellular domains of FtsL and DivIC did not interact in vitro (11, 28). Forced dimerization was obtained by fusion with artificial coiled-coil peptides k5 and e5 (35 residues long), which are known to form a heterodimer due to their complementarity of charge, with a nanomolar dissociation constant (29). The k5- and e5-coils were fused to the extracellular domain of FtsL and DivIC, to give rise to KL and EC fusion proteins, respectively. The constrained dimer (KL/EC) was shown to interact with the extracellular part of DivIB (DivIBext), yielding a soluble complex amenable to structural studies (11).The overall shape of the complex and its constituents was probed using small-angle x-ray scattering (SAXS)2 and small-angle neutron scattering (SANS). NMR was used to investigate the interface between the proteins by chemical shift mapping. The interaction was further investigated using surface plasmon resonance with truncated forms of the proteins. The complex of DivIB, FtsL, and DivIC is formed by the interaction of one face of the β-domain of DivIB with the C-terminal regions of FtsL and DivIC, at the tip of an elongated rod formed by the coiled-coil segments. The α-domain of DivIB and the coiled-coil regions of FtsL and DivIC remain free to interact with other proteins of the division apparatus.  相似文献   
1000.
The paralyzed zebrafish strain relaxed carries a null mutation for the skeletal muscle dihydropyridine receptor (DHPR) β1a subunit. Lack of β1a results in (i) reduced membrane expression of the pore forming DHPR α1S subunit, (ii) elimination of α1S charge movement, and (iii) impediment of arrangement of the DHPRs in groups of four (tetrads) opposing the ryanodine receptor (RyR1), a structural prerequisite for skeletal muscle-type excitation-contraction (EC) coupling. In this study we used relaxed larvae and isolated myotubes as expression systems to discriminate specific functions of β1a from rather general functions of β isoforms. Zebrafish and mammalian β1a subunits quantitatively restored α1S triad targeting and charge movement as well as intracellular Ca2+ release, allowed arrangement of DHPRs in tetrads, and most strikingly recovered a fully motile phenotype in relaxed larvae. Interestingly, the cardiac/neuronal β2a as the phylogenetically closest, and the ancestral housefly βM as the most distant isoform to β1a also completely recovered α1S triad expression and charge movement. However, both revealed drastically impaired intracellular Ca2+ transients and very limited tetrad formation compared with β1a. Consequently, larval motility was either only partially restored (β2a-injected larvae) or not restored at all (βM). Thus, our results indicate that triad expression and facilitation of 1,4-dihydropyridine receptor (DHPR) charge movement are common features of all tested β subunits, whereas the efficient arrangement of DHPRs in tetrads and thus intact DHPR-RyR1 coupling is only promoted by the β1a isoform. Consequently, we postulate a model that presents β1a as an allosteric modifier of α1S conformation enabling skeletal muscle-type EC coupling.Excitation-contraction (EC)3 coupling in skeletal muscle is critically dependent on the close interaction of two distinct Ca2+ channels. Membrane depolarizations of the myotube are sensed by the voltage-dependent 1,4-dihydropyridine receptor (DHPR) in the sarcolemma, leading to a rearrangement of charged amino acids (charge movement) in the transmembrane segments S4 of the pore-forming DHPR α1S subunit (1, 2). This conformational change induces via protein-protein interaction (3, 4) the opening of the sarcoplasmic type-1 ryanodine receptor (RyR1) without need of Ca2+ influx through the DHPR (5). The release of Ca2+ from the sarcoplasmic reticulum via RyR1 consequently induces muscle contraction. The protein-protein interaction mechanism between DHPR and RyR1 requires correct ultrastructural targeting of both channels. In Ca2+ release units (triads and peripheral couplings) of the skeletal muscle, groups of four DHPRs (tetrads) are coupled to every other RyR1 and hence are geometrically arranged following the RyR-specific orthogonal arrays (6).The skeletal muscle DHPR is a heteromultimeric protein complex, composed of the voltage-sensing and pore-forming α1S subunit and auxiliary subunits β1a, α2δ-1, and γ1 (7). While gene knock-out of the DHPR γ1 subunit (8, 9) and small interfering RNA knockdown of the DHPR α2δ-1 subunit (10-12) have indicated that neither subunit is essential for coupling of the DHPR with RyR1, the lack of the α1S or of the intracellular β1a subunit is incompatible with EC coupling and accordingly null model mice die perinatally due to asphyxia (13, 14). β subunits of voltage-gated Ca2+ channels were repeatedly shown to be responsible for the facilitation of α1 membrane insertion and to be potent modulators of α1 current kinetics and voltage dependence (15, 16). Whether the loss of EC coupling in β1-null mice was caused by decreased DHPR membrane expression or by the lack of a putative specific contribution of the β subunit to the skeletal muscle EC coupling apparatus (17, 18) was not clearly resolved. Recently, other β-functions were identified in skeletal muscle using the β1-null mutant zebrafish relaxed (19, 20). Like the β1-knock-out mouse (14) zebrafish relaxed is characterized by complete paralysis of skeletal muscle (21, 22). While β1-knock-out mouse pups die immediately after birth due to respiratory paralysis (14), larvae of relaxed are able to survive for several days because of oxygen and metabolite diffusion via the skin (23). Using highly differentiated myotubes that are easy to isolate from these larvae, the lack of EC coupling could be described by quantitative immunocytochemistry as a moderate ∼50% reduction of α1S membrane expression although α1S charge movement was nearly absent, and, most strikingly, as the complete lack of the arrangement of DHPRs in tetrads (19). Thus, in skeletal muscle the β subunit enables EC coupling by (i) enhancing α1S membrane targeting, (ii) facilitating α1S charge movement, and (iii) enabling the ultrastructural arrangement of DHPRs in tetrads.The question arises, which of these functions are specific for the skeletal muscle β1a and which ones are rather general properties of Ca2+ channel β subunits. Previous reconstitution studies made in the β1-null mouse system (24, 25) using different β subunit constructs (26) did not allow differentiation between β-induced enhancement of non-functional α1S membrane expression and the facilitation of α1S charge movement, due to the lack of information on α1S triad expression levels. Furthermore, the β-induced arrangement of DHPRs in tetrads was not detected as no ultrastructural information was obtained.In the present study, we established zebrafish mutant relaxed as an expression system to test different β subunits for their ability to restore skeletal muscle EC coupling. Using isolated myotubes for in vitro experiments (19, 27) and complete larvae for in vivo expression studies (28-31) and freeze-fracture electron microscopy, a clear differentiation between the major functional roles of β subunits was feasible in the zebrafish system. The cloned zebrafish β1a and a mammalian (rabbit) β1a were shown to completely restore all parameters of EC coupling when expressed in relaxed myotubes and larvae. However, the phylogenetically closest β subunit to β1a, the cardiac/neuronal isoform β2a from rat, as well as the ancestral βM isoform from the housefly (Musca domestica), could recover functional α1S membrane insertion, but led to very restricted tetrad formation when compared with β1a, and thus to impaired DHPR-RyR1 coupling. This impairment caused drastic changes in skeletal muscle function.The present study shows that the enhancement of functional α1S membrane expression is a common function of all the tested β subunits, from β1a to even the most distant βM, whereas the effective formation of tetrads and thus proper skeletal muscle EC coupling is an exclusive function of the skeletal muscle β1a subunit. In context with previous studies, our results suggest a model according to which β1a acts as an allosteric modifier of α1S conformation. Only in the presence of β1a, the α1S subunit is properly folded to allow RyR1 anchoring and thus skeletal muscle-type EC coupling.  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