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   231篇
  免费   7篇
化学   151篇
晶体学   1篇
力学   10篇
数学   8篇
物理学   68篇
  2023年   1篇
  2022年   1篇
  2019年   2篇
  2018年   4篇
  2017年   3篇
  2016年   3篇
  2015年   3篇
  2014年   4篇
  2013年   13篇
  2012年   4篇
  2011年   12篇
  2010年   10篇
  2009年   2篇
  2008年   26篇
  2007年   26篇
  2006年   14篇
  2005年   11篇
  2004年   17篇
  2003年   13篇
  2002年   7篇
  2001年   7篇
  2000年   6篇
  1999年   2篇
  1996年   1篇
  1994年   3篇
  1993年   2篇
  1990年   1篇
  1989年   3篇
  1988年   2篇
  1987年   7篇
  1986年   1篇
  1985年   3篇
  1983年   1篇
  1982年   1篇
  1981年   1篇
  1980年   2篇
  1979年   1篇
  1978年   3篇
  1977年   2篇
  1976年   1篇
  1974年   3篇
  1972年   1篇
  1969年   2篇
  1967年   2篇
  1965年   1篇
  1964年   1篇
  1906年   2篇
排序方式: 共有238条查询结果,搜索用时 15 毫秒
71.
Abstract— The anaerobic photobleaching of 6, 7-dimethyl-9-(2′-hydroxyethyl)-isoalloxazine, 9-(2′-hydroxyethyl)-isoalloxazine and 6,7-dimethyl-9-(formylmethyl)-isoalloxazine have been compared. Polarographic and spectrophotometric techniques were used for rate studies. The methyl groups on the aromatic ring at the 6 and 7 positions affects the product distributtion in these simple flavins. Formylmethylflavin is postulated to undergo a bimolecular photoredox reaction which is novel to flavin photochemistry. It was demonstrated that there is a common type of leucoflavin formed during anaerobic photobleaching of riboflavin, formylmethylflavin and hydroxyethylflavin. This leucocompound also formed during the photoreduction of formylmethylflavin but not during the photoreduction of the other flavins.  相似文献   
72.
The interactions between the stereoisomers of the chiral bis-intercalator [mu-C4(cpdppz)(2)-(phen)(4)Ru(2)](4+) and DNA reveal interesting dynamic discrimination properties. The two enantiomers Delta-Delta and Lambda-Lambda both form very strong complexes with calf thymus DNA with similar thermodynamic affinities. By contrast, they display considerable variations in their binding kinetics. The Delta-Delta enantiomer has higher affinity for calf thymus DNA than for [poly(dA-dT)](2), and the association kinetics of the dimer to DNA, as well as to polynucleotides, requires a multiexponential fitting function. The dissociation reaction, on the other hand, could be described by a single exponential for [poly(dA-dT)](2), whereas two exponentials were required for mixed-sequence DNA. To understand the key mechanistic steps of the reaction, the kinetics was studied at varied salt concentration for different choices of DNA and chirality of the threading complex. The enantiomers were found to have markedly different dissociation rates, the Lambda-Lambda enantiomer dissociating about an order of magnitude faster than the Delta-Delta enantiomer. Also, the salt dependence of the dissociation rate constants differed between the enantiomers, being stronger for the Lambda-Lambda enantiomer than for the Delta-Delta enantiomer. Since the dissociation reaction requires unthreading of bulky parts of the bis-intercalator through the DNA helix, a considerable conformational change of the DNA must be involved, possibly defining the rate-limiting step.  相似文献   
73.
Chaotic spin-wave solitons in magnetic film active feedback rings were observed for the first time. At some ring gain level, one observes the self-generation of a single spin-wave soliton pulse in the ring. When the pulse circulates in the ring, its amplitude varies chaotically with time. Numerical simulations based on a gain-loss nonlinear Schr?dinger equation reproduce the observed responses.  相似文献   
74.
Using 586 pb(-1) of e+ e- collision data at E(c.m.) = 4170 MeV, produced at the Cornell Electron Storage Ring collider and collected with the CLEO-c detector, we observe the process e+ e- → π+ π- h(c)(1P). We measure its cross section to be 15.6±2.3±1.9±3.0 pb, where the third error is due to the external uncertainty on the branching fraction of ψ(2S) → π0 h(c)(1P), which we use for normalization. We also find evidence for e+ e- → ηh(c)(1P) at 4170 MeV at the 3σ level and see hints of a rise in the e+ e- → π+ π- h(c)(1P) cross section at 4260 MeV.  相似文献   
75.
The crystal structure of the Δ,Δ enantiomer of the binuclear “light‐switch” ruthenium complex [μ‐(11,11′‐bidppz)(1,10‐phenanthroline)4 Ru2]4+ bound to the oligonucleotide d(CGTACG) shows that one dppz moiety of the dumbbell‐like compound inserts into the DNA stack through the extrusion of an AT base pair. The second dppz moiety recruits a neighboring DNA molecule, and the complex thus cross‐links two adjacent duplexes by bridging their major grooves.  相似文献   
76.
Two amidine‐substituted spiropyran derivatives have been characterized with respect to the DNA‐binding properties over a broad pH interval. The two derivatives differ in the number of positive charges. By varying the pH, the protonation state of the derivatives is also changed, allowing for additional variations in the charge distribution. We show that the closed spiro isomer does not bind for either of the two derivatives, whereas the open merocyanine forms bind both in the protonated and in the nonprotonated state, but with dramatically different binding constants. Flow‐oriented linear dichroism (LD) measurements also show that there are differences in the binding modes between the various forms. We rationalize these differences in terms of structure and charge distribution.  相似文献   
77.
The concept of Majorana and Dirac massive neutrinos is discussed for the case in which there are several types, or flavors, of leptons. Different varieties of Dirac neutrinos are possible, distinguised by their magnetic moments and anomalous interactions with probabilities proportional to (m/E)2.  相似文献   
78.
79.
The dumbbell shaped binuclear ruthenium complex DeltaDelta-P requires transiently melted DNA in order to thread through the DNA bases and intercalate DNA. Because such fluctuations are rare at room temperature, the binding rates are extremely low in bulk experiments. Here, single DNA molecule stretching is used to lower the barrier to DNA melting, resulting in direct mechanical manipulation of the barrier to DNA binding by the ligand. The rate of DNA threading depends exponentially on force, consistent with theoretical predictions. From the observed force dependence of the binding rate, we demonstrate that only one base pair must be transiently melted for DNA threading to occur.  相似文献   
80.
We recently reported that ruthenium complexes, with general structure [mu-bidppz(bipy)4Ru2](4+) (B) or [mu-bidppz(phen)4Ru2](4+) (P) (bidppz=11,11'-bi(dipyrido[3,2- a:2',3'-c]phenazinyl)), show extreme kinetic selectivity for long AT tracts over mixed-sequence calf thymus DNA (ct-DNA), a selectivity that also varies markedly with the size (between B and P) and sense of chirality of the complex. Earlier studies, exploiting the great increase in luminescence intensity when the compound intercalates, have yielded complex kinetics indicating the presence of both first- and second-order processes. Even with a homogeneous DNA sequence, such as poly(dAdT)2, the luminescence kinetics generally requires more than a single exponential for a satisfactory fit. We here reveal that at least part of the complexity is a result of the extreme sensitivity of the effective quantum yield of the complexes, so that the luminescence trajectories also reflect subtle variations in the environment and binding geometry that the complex is sampling on its path to its final binding site. By monitoring the rearrangement process using circular dichroism (CD), we show that threading of both enantiomers of B and P into poly(dAdT)2 is effectively a monoexponential process, as expected if the compounds are not affecting each other during the intercalation process. Thus, the complex luminescence trajectories may be explained by slow relaxations in the binding geometry (DNA conformation) and associated changes in the environment of the entering complexes. To further disentangle the intriguing features of the threading intercalation kinetics, and how they may depend on the flexibility and size of the ruthenium complexes, we have also designed and studied two new ruthenium complexes, [mu-dtpf(phen)4Ru2](4+) (F) (dtpf=4,5,9,12,16,17,21,25-octaaza-23 H-ditriphenyleno[2,3-b:2,3-h]fluorene), in which the bridging ligand is made totally rigid, and [mu-bidppz([12]aneS4) 2Ru2](4+) (S), which has less bulky, nonaromatic ancillary ligands. The threading of F into poly(dAdT)2, also found to be a monoexponential process, is about 3 times slower than for P, indicating that the flexibility of the bridging ligand is an important factor for the intercalation rate. Surprisingly, in contrast to all other compounds, S requires two exponentials to fit its binding kinetics as monitored by CD. Also surprisingly, in view of the smaller steric bulk, even the fastest phase is roughly 2 times slower for S than for B and P. Thus, not only the size of the ancillary ligand but also other properties that can influence the energy landscape of the threading path are rate-determining factors. With mixed-sequence ct-DNA, threading of B and that of P are both multiphasic processes when monitored with CD as well as with luminescence. The rate constants for threading into ct-DNA show much larger variations between complexes than for poly(dAdT)2, confirming earlier results based on luminescence data.  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