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   2128篇
  免费   64篇
  国内免费   16篇
化学   1480篇
晶体学   13篇
力学   44篇
数学   306篇
物理学   365篇
  2023年   11篇
  2022年   38篇
  2021年   75篇
  2020年   42篇
  2019年   48篇
  2018年   33篇
  2017年   29篇
  2016年   77篇
  2015年   57篇
  2014年   62篇
  2013年   128篇
  2012年   140篇
  2011年   126篇
  2010年   79篇
  2009年   87篇
  2008年   126篇
  2007年   128篇
  2006年   133篇
  2005年   117篇
  2004年   75篇
  2003年   58篇
  2002年   63篇
  2001年   20篇
  2000年   14篇
  1999年   17篇
  1998年   14篇
  1997年   17篇
  1996年   23篇
  1995年   18篇
  1994年   20篇
  1993年   23篇
  1992年   21篇
  1991年   15篇
  1990年   8篇
  1989年   8篇
  1988年   15篇
  1987年   11篇
  1985年   16篇
  1984年   11篇
  1983年   11篇
  1981年   8篇
  1980年   9篇
  1978年   9篇
  1977年   12篇
  1976年   7篇
  1975年   13篇
  1974年   12篇
  1973年   13篇
  1971年   11篇
  1908年   8篇
排序方式: 共有2208条查询结果,搜索用时 15 毫秒
991.
This report describes the synthesis, structural characterization, and polymerization behavior of a series of chromium(II) and chromium(III) complexes ligated by tris(2-pyridylmethyl)amine (TPA), including chromium(III) organometallic derivatives. For instance, the combination of TPA with CrCl(2) yields monomeric (TPA)CrCl(2) (1). A similar reaction of CrCl(2) with TPA, followed by chloride abstraction with NaBPh(4) or NaBAr(F)(4) (Ar(F) = 3,5-(CF(3))(2)C(6)H(3)), provides the weakly associated cationic dimers [(TPA)CrCl](2)[BPh(4)](2) (2A) and [(TPA)CrCl](2)[BAr(F)(4)](2) (2B), respectively. X-ray crystallographic analysis reveals that each chromium(II) center in 1, 2A, and 2B is a tetragonally elongated octahedron; such Jahn-Teller distortions are consistent with the observed high spin (S = 2) electronic configurations for these chromium(II) complexes. Likewise, reaction of CrCl(3)(THF)(3) with TPA, followed by anion metathesis with NaBPh(4) or NaBAr(F)(4), yields the monomeric, cationic chromium(III) complexes [(TPA)CrCl(2)][BPh(4)] (4A) and [(TPA)CrCl(2)][BAr(F)(4)] (4B), respectively. Treatment of 4A with methyl and phenyl Grignard reagents produces the cationic chromium(III) organometallic derivatives [(TPA)Cr(CH(3))(2)][BPh(4)] (5) and [(TPA)CrPh(2)][BPh(4)] (6), respectively. Similar reactions of 4A with organolithium reagents leads to intractable solids, presumably due to overreduction of the chromium(III) center. X-ray crystallographic analysis of 4A, 5, and 6 confirms that each possesses a largely undistorted octahedral chromium center, consistent with the observed S = (3)/(2) electronic ground states. Compounds 1, 2A, 2B, 4A, 4B, 5, and 6 are all active polymerization catalysts in the presence of methylalumoxane, producing low to moderate molecular weight high-density polyethylene.  相似文献   
992.
The authors recently published works in which the use of two novel equations for modeling the dispersive kinetics observed in various solid-state conversions are described. These equations are based on the assumptions of a ‘Maxwell-Boltzmann (M-B)-like’ distribution of activation energies and a first-order rate law. In the present work, it is shown that it may be possible to expand the approach to include mechanisms other than first-order, i.e. some of those commonly encountered in the field of thermal analysis, thus obtaining ‘dispersive versions’ of these kinetic models. The application of these dispersive kinetic models to the slightly sigmoidal, isothermal conversion-time (x-t) data of Rodante and co-workers for the degradation of the antibiotic, oxacillin, is described. This is done in an effort to test the limitations of the proposed dispersive models in describing kinetic data which is not clearly sigmoidal (i.e. as shown in previous works). Finally, it is demonstrated that, using graphical analysis, the typically sigmoidal x-t plots of first-order dispersive processes are the direct result of (asymmetric) activation energy distributions that are either ‘∩-shaped’ (for heterogeneous conversions) or ‘∪-shaped’ (for homogeneous conversions) in appearance, i.e. when the activation energy is plotted as a function of conversion. This finding lends support to the founding hypothesis of the authors’ approach for modeling dispersive kinetic processes: the existence of ‘M-B-like’ distributions of activation energies.  相似文献   
993.
Bacterial membrane vesicles retain the same sidedness as the membrane in the intact cell and catalyze active transport of many solutes by a respiration-dependent mechanism that does not involve the generation of utilization of ATP or other high-energy phosphate compounds. In E. coli vesicles, most of these transport systems are coupled to an electrochemical gradient of protons (deltamuH+, interior negative and alkaline) generated primarily by the oxidation of D-lactate or reduced phenazine methosulfate via a membrane-bound respiratory chain. Oxygen or, under appropriate conditions, fumarate or nitrate can function as terminal electron acceptors, and the site at which deltamuH+ is generated is located before cytochrome b1 in the respiratory chain. Certain (N-dansyl)aminoalkyl-beta-D-galactopyranosides (Dns-gal) and N(2-nitro-4-azidophenyl)aminoalkyl 1-thio-beta-D-galactopyranosides (APG) are competitive inhibitors of lactose transport but are not transported themselves. Various fluorescence techniques, direct binding assays, and photoinactivation studies demonstrate that the great bulk of the lac carrier protein (ca. 95%) does not bind ligand in the absence of energy-coupling. Upon generation of a deltamuH+ (interior negative and alkaline), binding of Dns-gal and APG-dependent photoinactivation are observed. The data indicate that energy is coupled to the initial step in the transport process, and suggest that the lac carrier protein may be negatively charged.  相似文献   
994.
We present theoretical studies based on first-principles density functional theory calculations on the mechanisms of chemical vapor deposition of Cu-hexafluoracetylacetonato-trimethylvinylsilane (Cu(hfac)(tmvs)) on tantalum surfaces. This process has been used in the past to grow copper films via a disproportionation reaction and was found to exhibit adhesion problems. We show that the Ta surfaces are highly reactive and that the organic ligands in a copper precursor would undergo spontaneous decomposition upon contact with the Ta substrates. This may lead to contamination of the metal surfaces caused by the formation of carbide, fluoride, oxide species, or other fragments of the copper precursor on the barrier layer. We propose a practical solution for these adhesion problems caused by the CVD process by passivating the metal surfaces with N(2) to reduce their activity toward the precursor. Our extensive first-principles molecular dynamics simulations under typical deposition conditions predict that, for properly passivated TaN surfaces, only the copper atoms are firmly adsorbed on the surface, with loose Cu-ligand bonds. The ligands are sufficiently stable on these passivated surfaces, remaining slightly above the surface due to the repulsion between the electron-rich N-layer and the electron-rich ligand groups, and subsequently liberated upon the disproportionation reaction.  相似文献   
995.
The proximity of the calcium/strontium binding site of the oxygen evolving complex (OEC) of photosystem II (PSII) to the paramagnetic Mn cluster is explored with (87)Sr three-pulse electron spin-echo envelope modulation (ESEEM) spectroscopy. CW-EPR spectra of Sr(2+)-substituted Ca(2+)-depleted PSII membranes show the modified g = 2 multiline EPR signal as previously reported. We performed three-pulse ESEEM on this modified multiline signal of the Mn cluster using natural abundance Sr and (87)Sr, respectively. Three-pulse ESEEM of the natural abundance Sr sample exhibits no detectable modulation by the 7% abundance (87)Sr. On the other hand, that of the (87)Sr enriched (93%) sample clearly reveals modulation arising from the I = (9)/(2) (87)Sr nucleus weakly magnetically coupled to the Mn cluster. Using a simple point dipole approximation for the electron spin, analysis of the (87)Sr ESEEM modulation depth via an analytic expression suggests a Mn-Ca (Sr) distance of 4.5 A. Simulation of three-pulse ESEEM with a numerical matrix diagonalization procedure gave good agreement with this analytical result. A more appropriate tetranuclear magnetic/structural model for the Mn cluster converts the 4.5 A point dipole distance to a 3.8-5.0 A range of distances. DFT calculations of (43)Ca and (87)Sr quadrupolar interactions on Ca (and Sr substituted) binding sites in various proteins suggest that the lack of the nuclear quadrupole induced splitting in the ESEEM spectrum of (87)Sr enriched PSII samples is related to a very high degree of symmetry of the ligands surrounding the Sr(2+) ion in the substituted Ca site. Numerical simulations show that moderate (87)Sr quadrupolar couplings decrease the envelope modulation relative to the zero quadrupole case, and therefore we consider that the 3.8-5.0 A range obtained without quadrupolar coupling included in the simulation represents an upper limit to the actual manganese-calcium distance. This (87)Sr pulsed EPR spectroscopy provides independent direct evidence that the calcium/strontium binding site is close to the Mn cluster in the OEC of PSII.  相似文献   
996.
The compound 1-phenyl-1,2-dicarba-closo-dodecaborane(12), 1-C(6)H(5)-1,2-closo-C(2)B(10)H(11) (1), has been synthesized and characterized by a complete assignment of its (11)B NMR spectrum via (11)B{(1)H}/(11)B{(1)H} (COSY), (1)H{(11)B(selective)} and (1)H{(11)B}/(1)H{(11)B} (COSY) spectroscopy. An electron- and X-ray diffraction investigation of 1, complemented by ab initio calculations, has been undertaken. The gas-phase electron-diffraction (GED) data can be fitted by several models describing conformations which differ in the position of the phenyl ring with respect to the carborane cage. Local symmetries ofC(2)(v)() and D(6)(h)() for the 1,2-C(2)B(10) and C(6) moieties, respectively, were adopted in the GED model in order to simplify the problem. In addition, constraints among the close-lying C-C and B-B bonds were employed. However, even though such simplifications led to satisfactory refinements (R(G) = 0.069-0.071), a unique, definitive solution could not be gained. The (C-C)(mean), (C-B)(mean) and (B-B)(mean) bond lengths,r(a), are ca. 1.44, 1.72, and 1.78 ?, respectively. The C(6) hexagon, with r(a)(C-C) = ca. 1.394 ?, either eclipses the C(1)-C(2) vector (overall C(s)() symmetry) or more or less eclipses the C(1)-B(4) cluster bond (overall C(1) symmetry). In contrast, in the solid at 199 K, the ring lies at a position intermediate between the two GED positions, as determined by X-ray crystallography [C(8)H(16)B(10), monoclinic P2(1)/a: a = 12.047(3) ?, b = 18.627(4) ?, c = 12.332(5) ?, beta = 110.09(4) degrees, Z = 8]. The C-B distances span the range 1.681(6)-1.743(5) ?, and B-B lengths lie between 1.756(6) and 1.795(6) ?. A similar conformation was found for the theoretical (RHF/6-31G level) structure which was fully optimized in C(1) symmetry. The r(e) distances are consistent with the dimensions derived in the experimental studies. IGLO calculations of the (11)B chemical shifts, in addition to SCF single-point energies of the GED structures, further support these observations.  相似文献   
997.
We present a simple model to describe the phase relaxation of the νs(XH) mode of a hydrogen-bonded species XH?Y in solution in an inert solvent. The major assumptions made are that the νσ(XH?Y) stretching mode of the hydrogen bond can be represented by the Ornstein—Uhlenbeck stochastic process, and that the νs(XH) phase coherence is lost because of the coupling between the two vibrational modes. The model can be analysed in terms of Kubo's theory of a randomly modulated oscillator. An expression is derived for the transition dipole moment autocorrelation function which characterises the line shape of the mid infrared νs(XH) absorption band, and various limiting cases of the formula are discussed. It is further shown that vibrational relaxation may be expected to influence the far infrared νσ(XH?Y) bandshape and the one-phonon neutron inelastic scattering cross section, and that the parameters required to characterise all three types of spectra are closely related. Experimental tests of the theory are reported in the following paper.  相似文献   
998.
The molecular structures of the three closo-carbaboranes, ortho-, meta- and para-C2B10H12, were experimentally determined using gas-phase electron diffraction (GED). All unique bond distances for ortho and meta carbaboranes were determined experimentally for the first time. For ortho-carbaborane (RG= 0.046), a model with C2v symmetry refined to give bond distances of 1.624(8) A for C-C, 1.093(8)A for C-H and 1.192(3)-1.196(3) A for B-H. For meta-carbaborane (RG= 0.040) a model with C2v symmetry refined to give a CC distance of 2.575(9) A. For para-carbaborane (RG= 0.062) a model with D5d symmetry refined to give a C-B bond distance of 1.698(3) A, B2-B3 of 1.785(1), B2-B7 of 1.774(4) and CC of 3.029(5)A. These GED structures are compared with geometries from other experimental diffraction methods (neutron, X-ray) and ab initio calculations.  相似文献   
999.
The FTIR spectra of uracil and thymine were studied at different concentrations in pure Ar matrices as well as in H2O or HCl doped Ar matrices. The spectral results suggest the presence of open dimers in associated uracil and thymine. The structure of heterodimers of uracils with H2O and HCl is discussed.  相似文献   
1000.
Applying a local Gauss–Bonnet formula for closed subanalytic sets to the complex analytic case, we obtain characterizations of the Euler obstruction of a complex analytic germ in terms of the Lipschitz–Killing curvatures and the Chern forms of its regular part. We also prove analogous results for the global Euler obstruction. As a corollary, we give a positive answer to a question of Fu on the Euler obstruction and the Gauss–Bonnet measure.  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