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   1514篇
  免费   108篇
  国内免费   73篇
  2024年   2篇
  2023年   11篇
  2022年   14篇
  2021年   23篇
  2020年   22篇
  2019年   25篇
  2018年   37篇
  2017年   28篇
  2016年   30篇
  2015年   34篇
  2014年   44篇
  2013年   91篇
  2012年   44篇
  2011年   37篇
  2010年   42篇
  2009年   49篇
  2008年   82篇
  2007年   62篇
  2006年   62篇
  2005年   77篇
  2004年   70篇
  2003年   75篇
  2002年   76篇
  2001年   49篇
  2000年   29篇
  1999年   31篇
  1998年   79篇
  1997年   65篇
  1996年   34篇
  1995年   33篇
  1994年   45篇
  1993年   31篇
  1992年   25篇
  1991年   21篇
  1990年   24篇
  1989年   23篇
  1988年   16篇
  1987年   17篇
  1986年   10篇
  1985年   21篇
  1984年   27篇
  1983年   19篇
  1982年   26篇
  1981年   12篇
  1980年   6篇
  1979年   7篇
  1978年   2篇
  1977年   2篇
  1974年   1篇
  1973年   2篇
排序方式: 共有1695条查询结果,搜索用时 15 毫秒
951.
《Cell》2021,184(25):6157-6173.e24
  1. Download : Download high-res image (145KB)
  2. Download : Download full-size image
  相似文献   
952.
Three series of model peptides containing histidine have been examined by 1H-n.m.r. and c.d. spectroscopy: X-His peptides with X = Gly, Ala, Leu; His-X peptides with X = Gly, Ala, Leu, Ser, Lys, Phe, Tyr; and Pro-His-X peptides with X = Gly; Ala; Leu; Val; Phe; Tyr, C.d. spectra were obtained for pH values between 1 and 11 to give titration curves [θ] vs. pH; 1H-n.m.r. spectra were recorded at four selected pH values corresponding to defined ionic species. 1H-n.m.r. spectra in Me2SO of the NH3+, Imid+, COO? ionic state (pH 4.5) were also obtained. The histidine side chain conformation in the various peptides and the changing ionic states is reflected in the 3Jαβ,β coupling constants, the Δδ ββ′ anisochrony values and the c.d. histidine chromophore contribution at 215 nm, and qualitative and semiquantitative correlations can be established between these parameters. Whereas the histidine side chain conformation is quite different in each of the three series, and varies with the ionic state and environment, it is practically identical for each peptide within a series: the nature of the X-residue does not exert any influence on the histidine side chain conformational behaviour. Thus, the classical rotamer distribution R I > R II > R III which is due to steric factors is usually observed unless specific intramolecular interactions such as hydrogen or ionic bonds override these.  相似文献   
953.
The diphosphinite ligand 9,9-(Ph2POCH2)2-fluorene (1) was reacted with group 10 metal dichlorides to form chelate complexes of formula [MCl2(1)] (MNi, 2; MPd, 3; MPt, 4) showing 8-membered metallocycles. Chloride abstraction from 3 with AgOTf afforded the dinuclear complex [M(μ-Cl)Pd(1)]2(OTf)2 (5), in which the ligand adopts a different conformation with respect of 3. In 5, the fluorene moiety and the phenyl groups display stabilizing interactions with the anion which is located close to the metal centre. With Fe(II), Co(II) and Zn(II) chlorides, the non-isolated intermediates [MCl2(1)] readily undergo oxidation to [MCl2(1ox)] (MFe, 6; MCo, 7; MZn, 8; 1ox = 9,9-(Ph2P(O)OCH2)2-fluorene) in which the diphosphinate ligand and the metal centre form 10-membered metallocycles. Complexes 6-8 are the first examples of structurally characterized diphosphinate metal chelates. The Zn(II) diphosphinite complex [ZnCl2(1)] (9) could be observed by NMR spectroscopy, along with the mixed phosphinite-phosphinate, mono-oxidized complex which is an intermediate in the formation of 8. Complex [ZnCl2(9.9-fluorene-dimethanol)(Ph2P(O)H)] (10) was also observed as hydrolysis product of 9. The X-ray molecular structures of 2, 3, 5.2OTf, 6, 7, 8 and 10 are reported.  相似文献   
954.
Mutations of the receptor tyrosine kinase KIT are linked to certain cancers such as gastrointestinal stromal tumors (GISTs). Biophysical, biochemical, and structural studies have provided insight into the molecular basis of resistance to the KIT inhibitors, imatinib and sunitinib. Here, solution‐phase hydrogen/deuterium exchange (HDX) and direct binding mass spectrometry experiments provide a link between static structure models and the dynamic equilibrium of the multiple states of KIT, supporting that sunitinib targets the autoinhibited conformation of WT‐KIT. The D816H mutation shifts the KIT conformational equilibrium toward the activated state. The V560D mutant exhibits two low energy conformations: one is more flexible and resembles the D816H mutant shifted toward the activated conformation, and the other is less flexible and resembles the wild‐type KIT in the autoinhibited conformation. This result correlates with the V560D mutant exhibiting a sensitivity to sunitinib that is less than for WT KIT but greater than for KIT D816H. These findings support the elucidation of the resistance mechanism for the KIT mutants.  相似文献   
955.
Protein secondary structure elements are arranged in distinct structural motifs such as four-α-helix bundle, 8α/8β TIM-barrel, Rossmann dinucleotide binding fold, assembly of a helical rod. Each structural motif is characterized by a particular type of helix-helix interactions. A unique pattern of contacts is formed by interacting helices of the structural motif. In each type of fold, edges of the helix surface, which participate in the formation of helix-helix contacts with preceding and following helices, differ. This work shows that circular arrangements of the four, eight, and sixteen α-helices, which are found in the four-α-helical motif, TIM-barrel 8α/8β fold, and helical rod of 16.3¯ helices per turn correspondingly, can be associated with the mutual positioning of the edges of the helix surfaces. Edges (i, i+1)−(i+1, i+2) of the helix surface are central for the interhelical contacts in a four-α-helix bundle. Edges (i, i+1)−(i+2, i+3) are involved in the assembly of four-α-helix subunits into helical rod of a tobacco mosaic virus and a three-helix fragment of a Rossmann fold. In 8α/8β TIM-barrel fold, edges (i, i+1)−(i+5, i+6) are involved in the octagon arrangement. Approximation of a cross section of each motif with a polygon (n-gon, n=4, 8, 16) shows that a good correlation exists between polygon interior angles and angles formed by the edges of helix surfaces.  相似文献   
956.
The synthesis and biological evaluation of penicillamine(6)-5-tert-butylproline(7)-oxytocin analogs and comparison with their proline(7)-oxytocin counterparts has led to the discovery of two potent oxytocin (OT) antagonists: [dPen(1),Pen(6)]-oxytocin (1, pA(2) = 8.22, EC(50) = 6.0 nM) and [dPen(1),Pen(6),5-tBuPro(7)]-oxytocin (2, pA(2) = 8.19, EC(50) = 6.5 nM). In an attempt to understand the conformational requirements for their biological activity, spectroscopic analyses of 1 and 2 were performed using (1)H NMR, laser Raman and CD techniques. In H(2)O, oxytocin analogs 1 and 2 exhibited cis-isomer populations of 7% and 35%, respectively. Measurement of the amide proton temperature coefficients revealed solvent shielded hydrogens for Gln(4) and Pen(6) in the major trans-conformer of 1 as well as for Gln(4) in the minor cis-conformer of 2. Few long-distance NOEs were observed, suggesting conformational averaging for analogs 1 and 2 in water; moreover, a lower barrier (16.6 +/- 0.2 kcal/mol) for isomerization of the amide N-terminal to 5-tBuPro(7) relative to OT was calculated from measuring the coalescence temperature of the Gly(9) backbone NH signals in the NMR spectra of 2. Observed bands in the Raman spectra of 1 and 2 correspond to C(beta)-S-S-C(beta) dihedral angles of +110-115 degrees and +/-90 degrees , respectively. In water, acetonitrile and methanol, the CD spectra for 1 exhibited a positive maximum around 236-239 nm; in trifluoroethanol, the spectra shifted and a negative maximum was observed at 240 nm. The CD spectra of 2 were unaffected by solvent changes and exhibited a negative maximum at 236-239 nm. The CD and Raman data both suggested that a conformation having a right-handed screw sense about the disulfide and a chi(CS-SC) dihedral angle value close to 115 degrees was favored for analog 1 in water, methanol and acetonitrile, but not trifluoroethanol, where a +/-90 degrees angle was favored. Analog 2 was more resilient to conformational change about the disulfide, and adopted a preferred disulfide geometry corresponding to a +/-90 degrees chi(CS-SC) dihedral angle. Monte Carlo conformational analysis of analogs 1 and 2 using distance restraints derived from NMR spectroscopy revealed two prominent conformational minima for analog 1 with disulfide geometries around +114 degrees and +116 degrees . Similar analysis of analog 2 revealed one conformational minimum with a disulfide geometry around +104 degrees . In sum, the conformation about the disulfide in [dPen(1),Pen(6)]-OT (1) was shown to be contingent on environment and in TFE, adopted a geometry similar to that of [dPen(1),Pen(6),5-tBuPro(7)]-OT (2) which appeared to be stabilized by hydrophobic interactions between the 5-tBuPro(7) (5R)-tert-butyl group, the Leu(8) isopropyl sidechain and the Pen(6)beta-methyl substituents. In light of the conformational rigidity of 2 about the disulfide bond, and the similar geometry adopted by 1 in TFE, a S-S dihedral angle close to +110 degrees may be a prerequisite for their binding at the receptor.  相似文献   
957.
Alkalinity refers to the substance that can react with acid in an aqueous system, which measures its capac-ity to neutralize hydrogen iron (H ). Alkalinity influ-ences greatly the stability and treatment ability of an-aerobic bioreactor. It has been well …  相似文献   
958.
cis 5-tert-butyl-L-proline (Cbp) was prepared rapidly and efficiently by the addition of low-valent tert-butyl cuprate to an aminal derived from proline. [Cbp(2), D-Leu(5)]-OP was then synthesized, showing a predominant cis peptide bond conformation, as confirmed by NMR.  相似文献   
959.
The structure of the complex mixture of the core oligosaccharide components of the lipooligosaccharide (LOS) fraction of Agrobacterium tumefaciens strain TT111 was determined directly on the deacetylated products by means of spectroscopical methods. The rhamnan oligosaccharide elongating the inner Kdo residue shares structural features with other polysaccharides from well-known plant pathogenic bacteria. Its conformation was determined through extensive molecular dynamic (MD) analysis and presents an epitope similar to that recognized from the plant defense system.  相似文献   
960.
We report here the construction of a mutant version of Escherichia coli alkaline phosphatase (AP) in which the active site Ser was replaced by Thr (S102T), in order to investigate whether the enzyme can utilize Thr as the nucleophile and whether the rates of the critical steps in the mechanism are altered by the substitution. The mutant AP with Thr at position 102 exhibited an approximately 4000-fold decrease in k(cat) along with a small decrease in Km. The decrease in catalytic efficiency of approximately 2000-fold was a much smaller drop than that observed when Ala or Gly were substituted at position 102. The mechanism by which Thr can substitute for Ser in AP was further investigated by determining the X-ray structure of the S102T enzyme in the presence of the Pi (S102T_Pi), and after soaking the crystals with substrate (S102T_sub). In the S102T_Pi structure, the Pi was coordinated differently with its position shifted by 1.3 A compared to the structure of the wild-type enzyme in the presence of Pi. In the S102T_sub structure, a covalent Thr-Pi intermediate was observed, instead of the expected bound substrate. The stereochemistry of the phosphorus in the S102T_sub structure was inverted compared to the stereochemistry in the wild-type structure, as would be expected after the first step of a double in-line displacement mechanism. We conclude that the S102T mutation resulted in a shift in the rate-determining step in the mechanism allowing us to trap the covalent intermediate of the reaction in the crystal.  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