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   718篇
  免费   44篇
  国内免费   8篇
妇产科学   1篇
基础医学   60篇
口腔科学   2篇
临床医学   30篇
内科学   444篇
神经病学   8篇
特种医学   10篇
外科学   12篇
综合类   40篇
预防医学   11篇
药学   141篇
中国医学   8篇
肿瘤学   3篇
  2023年   1篇
  2022年   20篇
  2021年   40篇
  2020年   15篇
  2019年   11篇
  2018年   15篇
  2017年   11篇
  2016年   23篇
  2015年   44篇
  2014年   74篇
  2013年   81篇
  2012年   86篇
  2011年   66篇
  2010年   52篇
  2009年   44篇
  2008年   59篇
  2007年   16篇
  2006年   14篇
  2005年   24篇
  2004年   9篇
  2003年   3篇
  2002年   5篇
  2001年   2篇
  2000年   3篇
  1998年   8篇
  1996年   7篇
  1995年   3篇
  1994年   2篇
  1993年   2篇
  1992年   4篇
  1991年   5篇
  1990年   4篇
  1989年   4篇
  1988年   1篇
  1987年   2篇
  1986年   4篇
  1985年   2篇
  1982年   1篇
  1980年   1篇
  1979年   1篇
  1978年   1篇
排序方式: 共有770条查询结果,搜索用时 15 毫秒
1.
A dual, double antigen, time-resolved fluorescence immunoassay (DELFIA) for the simultaneous detection and quantitation of diphtheria (D) and tetanus (T) antibodies in sera has been developed. In the double antigen format one arm of the antibody binds to antigen coated microtitre wells and the other arm binds to labelled antigen to provide a fluorescent signal. This assay was found to be functionally specific for IgG antibodies and showed a good correlation with established toxin neutralization assays. Furthermore, the double antigen set-up was species independent, permitting the direct use of existing international references of animal origin to measure protective antibody levels in humans in international units (IU/ml). The detection limit corresponded to 0.0003 IU/ml with Eu3+-labelled toxoids and to 0.0035 IU/ml using Sm3+-labelled toxoids. The assay was fast with a high capacity making it a suitable method for serological surveillance studies.  相似文献   
2.
王书军  高文远  贾伟  肖培根 《中草药》2005,36(8):1216-1220
目的为了能够更加充分了解贝母属药用植物,通过各种分析方法对两种贝母——浙贝母和平贝母中所包含的淀粉的物理化学性质进行了研究。方法采用X射线衍射,扫描电子显微镜(SEM)以及热分析(TGA)的方法对两种贝母中淀粉的性质进行了比较。结果通过研究发现,两种贝母淀粉的晶体类型都为典型的B型,这与马铃薯淀粉的晶体类型是一致的。浙贝母和平贝母淀粉的结晶度分别为29.9%和20.1%,而马铃薯淀粉的结晶度为44.9%。从两种贝母淀粉的结晶度可以看出,平贝母淀粉中直链淀粉的量要高于浙贝母淀粉中直链淀粉的量。两种贝母淀粉的颗粒尺寸为5~40μm,而且他们都小于马铃薯淀粉的颗粒尺寸。两种贝母淀粉颗粒的形状是圆形的和椭圆形的。热稳定性表明由于植物来源的不同导致淀粉颗粒结构不同,从而热稳定性存在明显的差异。结论两种贝母淀粉由于来源不同,物理化学性质存在明显的差异。  相似文献   
3.
PurposeTo investigate and compare venous sac and feeding artery embolization (VFE) with feeding artery embolization (FAE) alone for treatment of pulmonary arteriovenous malformations (PAVMs), based on difference in outcomes in decrease of the size of the draining vein.Materials and MethodsTwenty-six patients (7 male and 19 female; median age [interquartile range], 58 years [46–65 years]) with 42 simple PAVMs treated with coil embolization between August 2005 and December 2018 were retrospectively evaluated. Twenty PAVMs were treated with FAE early in the study period and compared with 22 PAVMs treated with VFE later in the study period. Follow-up computed tomography images obtained 8–20 months after embolotherapy were used for outcome analysis. Data related to patient demographics; follow-up period; baseline diameters of the feeding artery, venous sac, and draining vein; draining vein diameter after treatment; and decrease in the size of the draining vein, including the number reaching a threshold of 70% decrease, were compared between the 2 groups.ResultsThe draining vein decreased in size by a median of 46.4% in the FAE group and 66.3% in the VFE group, and the difference between the 2 groups was statistically significant (P = .009). There were no significant differences in the other parameters.ConclusionsVFE leads to a greater decrease in the size of the draining vein than FAE, suggesting that VFE results in more complete occlusion than FAE for treatment of PAVMs.  相似文献   
4.
5.
Rubber oxygenase A (RoxA) is one of only two known enzymes able to catalyze the oxidative cleavage of latex for biodegradation. RoxA acts as a processive dioxygenase to yield the predominant product 12-oxo-4,8-dimethyl-trideca-4,8-diene-1-al (ODTD), a tri-isoprene unit. Here we present a structural analysis of RoxA from Xanthomonas sp. strain 35Y at a resolution of 1.8 Å. The enzyme is a 75-kDa diheme c-type cytochrome with an unusually low degree of secondary structure. Analysis of the heme group arrangement and peptide chain topology of RoxA confirmed a distant kinship with diheme peroxidases of the CcpA family, but the proteins are functionally distinct, and the extracellular RoxA has evolved to have twice the molecular mass by successively accumulating extensions of peripheral loops. RoxA incorporates both oxygen atoms of its cosubstrate dioxygen into the rubber cleavage product ODTD, and we show that RoxA is isolated with O2 stably bound to the active site heme iron. Activation and cleavage of O2 require binding of polyisoprene, and thus the substrate needs to use hydrophobic access channels to reach the deeply buried active site of RoxA. The location and nature of these channels support a processive mechanism of latex cleavage.  相似文献   
6.
RAF kinase functions in the mitogen-activated protein kinase (MAPK) pathway to transmit growth signals to the downstream kinases MEK and ERK. Activation of RAF catalytic activity is facilitated by a regulatory complex comprising the proteins CNK (Connector enhancer of KSR), HYP (Hyphen), and KSR (Kinase Suppressor of Ras). The sterile alpha-motif (SAM) domain found in both CNK and HYP plays an essential role in complex formation. Here, we have determined the x-ray crystal structure of the SAM domain of CNK in complex with the SAM domain of HYP. The structure reveals a single-junction SAM domain dimer of 1:1 stoichiometry in which the binding mode is a variation of polymeric SAM domain interactions. Through in vitro and in vivo mutational analyses, we show that the specific mode of dimerization revealed by the crystal structure is essential for RAF signaling and facilitates the recruitment of KSR to form the CNK/HYP/KSR regulatory complex. We present two docking-site models to account for how SAM domain dimerization might influence the formation of a higher-order CNK/HYP/KSR complex.  相似文献   
7.
Chloride ion–pumping rhodopsin (ClR) in some marine bacteria utilizes light energy to actively transport Cl into cells. How the ClR initiates the transport is elusive. Here, we show the dynamics of ion transport observed with time-resolved serial femtosecond (fs) crystallography using the Linac Coherent Light Source. X-ray pulses captured structural changes in ClR upon flash illumination with a 550 nm fs-pumping laser. High-resolution structures for five time points (dark to 100 ps after flashing) reveal complex and coordinated dynamics comprising retinal isomerization, water molecule rearrangement, and conformational changes of various residues. Combining data from time-resolved spectroscopy experiments and molecular dynamics simulations, this study reveals that the chloride ion close to the Schiff base undergoes a dissociation–diffusion process upon light-triggered retinal isomerization.

Chloride ion (Cl) concentration in some bacterial cells is regulated by rhodopsin proteins, generally known as halorhodopsin, or hR. These proteins use light energy to pump Cl into cells (1, 2). Light is harvested by a molecule of retinal, covalently linked to an essential lysine residue in the seventh transmembrane helix of GPCR-like (G protein–coupled receptor) proteins. Light activation causes retinal to isomerize from the all-trans to the 13-cis configuration. This change triggers subsequent conformational changes throughout the rhodopsin molecule and releases chloride into the cytoplasm. Retinal thermally relaxes to the all-trans configuration within milliseconds and is then ready for the next photocycle. Cl ions are transported from the extracellular (EC) side to the cytoplasmic (CP) side during each photocycle (3, 4).Light-driven ion-pumping rhodopsin can be used to develop artificial solar energy harvesting and optogenetics (58), but the molecular mechanism must be understood in detail for such applications. Despite the importance of hR, our current experimental data concerning the structure and dynamics of the protein remain very limited. A related protein, proton (H+)-pumping bacteriorhodopsin (bR) discovered in the early 1970s, has been extensively studied by multiple methods, including time-resolved spectroscopy, crystallography, mutagenesis, and computer simulation (912). In particular, recent studies using time-resolved serial femtosecond crystallography (TR-SFX) methods performed at X-ray free-electron laser (XFEL) facilities allow three-dimensional (3D) visualization of retinal isomerization and associated local conformational changes. These changes are accompanied by movement of protons from a donor aspartate group to an acceptor aspartate (1315). However, the central component of this process, the transported H+, is difficult to observe by X-ray crystallography and could not be directly traced in bR TR-SFX studies. Recently, a breakthrough was reported in a study on the sodium-pumping rhodopsin KR2 (K. eikastus rhodopsin 2), in which electron density signals of Na+ uptake were observed at Δt = 1 ms after laser illumination (16).Cl, a strong X-ray scatterer, can be directly observed from electron density maps. These maps provide first-hand information on the movement of ions as being transported within short timescales after light activation. Furthermore, hR and bR presumably share a common molecular mechanism despite transporting ions in opposite directions. A close relationship is strongly implied by the interconversion of the function of two rhodopsins. Outward H+-pumping bR can be converted to an inward Cl pump by changing a single residue (D85T) (17), while hR from the cyanobacterium, Mastigocladopsis repens, is reported to pump protons after a single mutation (T74D) (18). The chloride pump can therefore serve as a system analogous to the proton transporter and provide valuable information that is difficult to obtain directly from bR.In this study, we focus on chloride ion–pumping rhodopsin (ClR) from the marine flavobacterium Nonlabens marinus S1-08T (19). The conserved DTD motif (Asp85-Thr89-Asp96) of the bR family, residues 85, 89, and 96, is replaced by an NTQ motif (Asn98- Thr102-Gln109) in ClR (Fig. 1). The sequence identity of ClR and canonical bR from Halobacterium salinarum is only 27%, but the two proteins, nevertheless, have highly similar structures, including the disposition of the retinal chromophore. ClR structures at cryogenic and room temperatures clearly reveal an architecture composed of seven transmembrane helices (TM A to G) (2, 20, 21). The retinal is covalently linked to the Nζ atom of the Lys235 located on TM-G. Anomalous diffraction signals of the Br identify a stable binding site near the protonated Schiff base (PSB) and a plausible exit site on the CP side (Fig. 1A). Buried water molecules and locations of cavities inside ClR suggest a pathway for Cl uptake on the EC side, but the molecular mechanism for light-triggered Cl pumping remains obscure. Upon light activation, the Cl tightly held near the PSB must break free from its hydrogen bonding network (Fig. 1B). It then passes through a hydrophobic region to reach the CP side (Fig. 1C). Crystal structures of ClR were previously determined with crystals under continuous illumination of visible laser light. Intriguingly, these steady-state models revealed unexpected movement of the retinal, without indication of photo-isomerization (22). Steady-state measurements, which show averages of mixed states, are thus of limited use in deciphering the molecular mechanism of light-driven Cl pumping.Open in a separate windowFig. 1.Structure of ClR and a plausible pathway of Cl transport. (A) Cross-sections of ClR with the backbone structure shown in cartoon representation. Transmembrane helices are marked using letters A through G, and the C-terminal helix H in the cytoplasm is also indicated. Surfaces are clipped to show the cross-section colored in yellow and the model being sliced and then opened about the axis near the helix E. Water molecules and Cl ions are shown as red- and green-colored spheres. Blue curves indicate the path of ion entering ClR and the principal pumping direction after passing retinal. (B) Key residues near the Cl ion and retinal, together with the NTQ motif shown in stick representation. (C) Residues that form a hydrophobic region between the retinal and the cytoplasm are highlighted in ball-and-stick representation. The red arrow points to a major barrier that Cl needs to overcome. ClR backbone is shown in cartoon representation, with residues colored based on hydrophobicity (the blue to red spectrum corresponds to the hydrophobicity scale from hydrophilic to hydrophobic).  相似文献   
8.
Plasmodium falciparum and Toxoplasma gondii are widely studied parasites in phylum Apicomplexa and the etiological agents of severe human malaria and toxoplasmosis, respectively. These intracellular pathogens have evolved a sophisticated invasion strategy that relies on delivery of proteins into the host cell, where parasite-derived rhoptry neck protein 2 (RON2) family members localize to the host outer membrane and serve as ligands for apical membrane antigen (AMA) family surface proteins displayed on the parasite. Recently, we showed that T. gondii harbors a novel AMA designated as TgAMA4 that shows extreme sequence divergence from all characterized AMA family members. Here we show that sporozoite-expressed TgAMA4 clusters in a distinct phylogenetic clade with Plasmodium merozoite apical erythrocyte-binding ligand (MAEBL) proteins and forms a high-affinity, functional complex with its coevolved partner, TgRON2L1. High-resolution crystal structures of TgAMA4 in the apo and TgRON2L1-bound forms complemented with alanine scanning mutagenesis data reveal an unexpected architecture and assembly mechanism relative to previously characterized AMA–RON2 complexes. Principally, TgAMA4 lacks both a deep surface groove and a key surface loop that have been established to govern RON2 ligand binding selectivity in other AMAs. Our study reveals a previously underappreciated level of molecular diversity at the parasite–host-cell interface and offers intriguing insight into the adaptation strategies underlying sporozoite invasion. Moreover, our data offer the potential for improved design of neutralizing therapeutics targeting a broad range of AMA–RON2 pairs and apicomplexan invasive stages.Phylum Apicomplexa comprises >5,000 parasitic protozoan species, many of which cause devastating diseases on a global scale. Two of the most prevalent species are Toxoplasma gondii and Plasmodium falciparum, the causative agents of toxoplasmosis and severe human malaria, respectively (1, 2). The obligate intracellular apicomplexan parasites lead complex and diverse lifestyles that require invasion of many different cell types. Despite this diversity of target host cells, most apicomplexans maintain a generally conserved mechanism for active invasion (3). The parasite initially glides over the surface of a host cell and then reorients to place its apical end in close contact to the host-cell membrane. After this initial attachment, a circumferential ring of adhesion (termed the moving or tight junction) is formed, through which the parasite actively propels itself while concurrently depressing the host-cell membrane to create a nascent protective vacuole (4).Formation of the moving junction relies on a pair of highly conserved parasite proteins: rhoptry neck protein 2 (RON2) and apical membrane antigen 1 (AMA1). Initially, parasites discharge RON2 into the host cell membrane where an extracellular portion (domain 3; D3) serves as a ligand for AMA1 displayed on the parasite surface (58). Intriguingly, recent studies have shown that the AMA1–RON2 complex is an attractive target for therapeutic intervention (912). The importance of the AMA1–RON2 pairing is also reflected in the observation that many apicomplexan parasites encode functional paralogs that are generally expressed in a stage-specific manner (1315). We recently showed that, in addition to AMA1 and RON2, T. gondii harbors three additional AMA paralogs and two additional RON2 paralogs (14, 15): TgAMA2 forms a functional invasion complex with TgRON2 (15), TgAMA3 (also annotated as SporoAMA1) selectively coordinates TgRON2L2 (14), and TgAMA4 binds TgRON2L1 (15). Despite substantial sequence divergence, structural characterization of all AMA–RON2D3 complexes solved to date [TgAMA1–TgRON2D3 (16), PfAMA1–PfRON2D3 (17), and TgAMA3–TgRON2L2D3 (14)] reveal a largely conserved architecture and binding paradigm. Intriguingly, however, sequence analysis indicates that TgAMA4 and TgRON2L1 are likely to adopt substantially divergent structures with an atypical assembly mechanism.To investigate the functional implications of the AMA4–RON2L1 complex in T. gondii, we first established that TgAMA4 is part of a highly divergent AMA clade that includes the functionally important malaria vaccine candidate Plasmodium merozoite apical erythrocyte-binding ligand (MAEBL) (1820) and that TgRON2L1 displays a similar divergence consistent with coevolution of receptor and ligand. We then show that TgAMA4 and TgRON2L1 form a high-affinity binary complex and probe its overall architecture and underlying mechanism of assembly using crystal structures of TgAMA4 in the apo and TgRON2L1D3 bound forms. Finally, we show proof of principle that TgAMA4 and TgRON2L1 form a functional pairing capable of supporting host-cell invasion. Collectively, our study reveals exceptional molecular diversity at the parasite–host-cell interface that we discuss in the context of the unique invasion barriers encountered by the sporozoite.  相似文献   
9.
10.
Improved understanding of the relationship among structure, dynamics, and function for the enzyme phenylalanine hydroxylase (PAH) can lead to needed new therapies for phenylketonuria, the most common inborn error of amino acid metabolism. PAH is a multidomain homo-multimeric protein whose conformation and multimerization properties respond to allosteric activation by the substrate phenylalanine (Phe); the allosteric regulation is necessary to maintain Phe below neurotoxic levels. A recently introduced model for allosteric regulation of PAH involves major domain motions and architecturally distinct PAH tetramers [Jaffe EK, Stith L, Lawrence SH, Andrake M, Dunbrack RL, Jr (2013) Arch Biochem Biophys 530(2):73–82]. Herein, we present, to our knowledge, the first X-ray crystal structure for a full-length mammalian (rat) PAH in an autoinhibited conformation. Chromatographic isolation of a monodisperse tetrameric PAH, in the absence of Phe, facilitated determination of the 2.9 Å crystal structure. The structure of full-length PAH supersedes a composite homology model that had been used extensively to rationalize phenylketonuria genotype–phenotype relationships. Small-angle X-ray scattering (SAXS) confirms that this tetramer, which dominates in the absence of Phe, is different from a Phe-stabilized allosterically activated PAH tetramer. The lack of structural detail for activated PAH remains a barrier to complete understanding of phenylketonuria genotype–phenotype relationships. Nevertheless, the use of SAXS and X-ray crystallography together to inspect PAH structure provides, to our knowledge, the first complete view of the enzyme in a tetrameric form that was not possible with prior partial crystal structures, and facilitates interpretation of a wealth of biochemical and structural data that was hitherto impossible to evaluate.Mammalian phenylalanine hydroxylase (PAH) (EC 1.14.16.1) is a multidomain homo-multimeric protein whose dysfunction causes the most common inborn error in amino acid metabolism, phenylketonuria (PKU), and milder forms of hyperphenylalaninemia (OMIM 261600) (1). PAH catalyzes the hydroxylation of phenylalanine (Phe) to tyrosine, using nonheme iron and the cosubstrates tetrahydrobiopterin and molecular oxygen (2, 3). A detailed kinetic mechanism has recently been derived from elegant single-turnover studies (4). PAH activity must be carefully regulated, because although Phe is an essential amino acid, high Phe levels are neurotoxic. Thus, Phe allosterically activates PAH by binding to a regulatory domain. Phosphorylation at Ser16 potentiates the effects of Phe, with phosphorylated PAH achieving full activation at lower Phe concentrations than the unphosphorylated protein (5, 6). Allosteric activation by Phe is accompanied by a major conformational change, as evidenced by changes in protein fluorescence and proteolytic susceptibility, and by stabilization of a tetrameric conformer (3).There are >500 disease-associated missense variants of human PAH; the amino acid substitutions are distributed throughout the 452-residue protein and among all its domains (Fig. 1A) (79). Of those disease-associated variants that have been studied in vitro (e.g., ref. 10), some confound the allosteric response, and some are interpreted as structurally unstable. We also suggest that the activities of some disease-associated variants may be dysregulated by an altered equilibrium among conformers having different intrinsic levels of activity, arguing by analogy to the enzyme porphobilinogen synthase (PBGS) and its porphyria-associated variants (11). Consistent with this notion, we have recently established that PAH can assemble into architecturally distinct tetrameric conformers (12), and propose that these conformers differ in activity due to differences in active-site access. This idea has important implications for drug discovery, as it implies that small molecules could potentially modulate the conformational equilibrium of PAH, as has already been demonstrated for PBGS (e.g., ref. 13). Deciphering the relationship among PAH structure, dynamics, and function is a necessary first step in testing this hypothesis.Open in a separate windowFig. 1.The structure of PAH. (A) The annotated domain structure of mammalian PAH. (B) The 2.9 Å PAH crystal structure in orthogonal views, colored as in part A, subunit A is shown in ribbons; subunit B is as a Cα trace; subunit C is in sticks; and subunit D is in transparent spheres. In cyan, the subunits are labeled near the catalytic domain (Top); in red, they are labeled near the regulatory domain (Bottom). The dotted black circle illustrates the autoregulatory domain partially occluding the enzyme active site (iron, in orange sphere). (C) Comparison of the subunit structures of full-length PAH and those of the composite homology model; the subunit overlay aligns residues 144–410. The four subunits of the full-length PAH structure (the diagonal pairs of subunits are illustrated using either black or white) are aligned with the two subunits of 2PAH (cyan) and the one subunit of 1PHZ (orange). The catalytic domain is in spheres, the regulatory domain is in ribbons, and the multimerization domain is as a Cα trace. The arrow denotes where the ACT domain and one helix of 2PAH conflict.Numerous crystal structures are known for one- and two-domain constructs of mammalian PAH (14).

Table S1.

Mammalian PAH structures available in the PDB (August 2015)
Open in a separate windowEntries are colored by configuration of the active-site lid (approximately residues 130–150): “open,” “closed,” and “disordered.”*C, catalytic domain, M, multimerization domain; R, regulatory domain.wwPDB Chemical Component Dictionary (49).Disease-associated single-residue substituted variant (A313T).§Phosphorylated at Ser16.  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