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   173篇
  免费   3篇
  国内免费   2篇
化学   134篇
力学   6篇
数学   8篇
物理学   30篇
  2022年   2篇
  2021年   3篇
  2018年   2篇
  2017年   2篇
  2015年   6篇
  2013年   6篇
  2012年   8篇
  2011年   13篇
  2010年   7篇
  2009年   3篇
  2008年   15篇
  2007年   9篇
  2006年   8篇
  2005年   11篇
  2004年   7篇
  2003年   9篇
  2002年   8篇
  2001年   6篇
  2000年   4篇
  1999年   1篇
  1998年   4篇
  1997年   4篇
  1996年   4篇
  1995年   3篇
  1994年   3篇
  1993年   4篇
  1992年   4篇
  1991年   1篇
  1990年   4篇
  1989年   1篇
  1988年   1篇
  1985年   1篇
  1983年   1篇
  1981年   3篇
  1980年   1篇
  1978年   1篇
  1977年   1篇
  1975年   1篇
  1974年   1篇
  1973年   1篇
  1971年   1篇
  1943年   2篇
  1934年   1篇
排序方式: 共有178条查询结果,搜索用时 15 毫秒
61.
Crystallization of Na(2)VOP(2)O(7) from its aqueous solution results in formation of a one-dimensional inorganic polymer {Na(2)VO(H(2)O)P(2)O(7)·7H(2)O}(n) (1). When this polymer is dehydrated at elevated temperatures this polymer undergoes a phase transition to form the two-dimensional framework β-Na(2)VOP(2)O(7), which although previously reported had been difficult to access. Exchanging lithium for sodium via ion-exchange chromatography results in formation of a discrete, cyclic, tetramer species, Li(8)[VOP(2)O(7)(H(2)O)·4H(2)O](4) (2). Isolation of crystalline β-Li(2)VOP(2)O(7) using a dehydration procedure analogous to the one employed for the sodium derivative was unsuccessful. In contrast, we show that β-K(2)VOP(2)O(7) can be obtained from the amorphous phase K(2)VOP(2)O(7)·nH(2)O (n = 0-7) upon thermal dehydration.  相似文献   
62.
63.
Facile methoxymethylation of N(2)-derived nitride NMo(N[(t)Bu]Ar)(3) provided the imido cation [MeOCH(2)NMo(N[(t)Bu]Ar)(3)](+) as its triflate salt in 88% yield. Treatment of the latter with LiN(SiMe(3))(2) provided blue methoxyketimide complex MeO(H)CNMo(N[(t)Bu]Ar)(3) in 95% yield. Conversion of the latter to the terminal cyanide complex NCMo(N[(t)Bu]Ar)(3), which was the subject of a single-crystal X-ray diffraction study, was accomplished in 51% yield upon treatment with a combination of SnCl(2) and Me(2)NSiMe(3).  相似文献   
64.
65.
66.
Several improvements have been made to the gradient algorithms commonly used to optimize equilibrium and transition-state geometries at the semiempirical level. A gradient algorithm derived from a combination of a variable metric method (Davidon–Fletcher–Powell/Broyden–Fletcher–Goldfarb–Shanno) and Pulay's direct inversion in the iterative subspace method for geometry optimization (GDIIS) is compared with the variable metric method combined with an accurate linear search algorithm. The latter method is used routinely in the standard semiempirical program packages, MNDO, MOPAC, and AMPAC. The combined variable metric and GDIIS algorithm is also compared with GDIIS which uses a static metric. The performance of these algorithms is examined for a wide range of systems with respect to both choice of coordinate system (for cyclic molecules) and guess for the initial Hessian. The results show that the GDIIS method is up to ca. 40% more efficient than the variable metric combined with accurate line search algorithm: however, the exact savings vary depending on the coordinate system and initial Hessian. For noncyclic systems, variable-metric GDIIS is usually equal or superior to static-metric GDIIS, and consistently performs ca. 30% more efficiently than the variable metric combined with accurate line search algorithm. For the optimization of cyclic molecules, an improved estimate of the initial Hessian has increased the efficiency by at least a factor of two. Greater efficiencies (usually >40%) are also obtained when static-metric GDIIS is used to refine the geometry after the initial application of a transition-state search based on the variable metric combined with line search algorithm. On the basis of these results, we recommend several changes to the algorithms as currently implemented in the standard semiempirical program packages.  相似文献   
67.
68.
Electromotive-force (emf) measurements of cells containing solutions of hydrochloric acid and neodymium chloride were reported at constant total ionic strengths (I) of 0.01, 0.025, 0.05, 0.1, 0.25, 0.5, 1.0, and 1.5 mol-kg−1 at 11 temperatures ranging from 5 to 55 C, and at I = 2.0 mol-kg−1 at 25 C. Hydrogen and silver–silver chloride electrodes were used in these cells. Results from the emf measurements, the mean molal activity coefficients of HCl in HCl + NdCl3 + H2O mixtures, as well as the Harned interaction coefficients using Harned's rule are reported in the preceding article in this issue. The ion-interaction model of Pitzer is applied here for the evaluation of the Pitzer mixing coefficients, SθH,Nd and ψH,Cl,Nd, as well as the linear representation of the temperature derivatives of ∂SθH,Nd /∂ T and ∂ψH,Nd,Cl/∂T. The activity coefficients at several ionic strength fractions y of NdCl3 are given at 25 C. The results are interpreted in terms of ionic interactions.  相似文献   
69.
Treatment of IU(DME)(NC[(t)Bu]Mes)(3) (2-I-DME) with 4 equiv of KC(8) and 0.5 equiv of naphthalene in DME allowed the isolation of a naphthalene-bridged compound, K(2)(mu-eta(6),eta(6)-C(10)H(8))[U(NC[(t)Bu]Mes)(3)](2) (K(2)-2(2)-mu-C(10)H(8)), in 60% yield as a dark brown powder. The twelve U-C distances are rather short, varying from 2.565(11) to 2.749(10) A. Treatment of M(2)-2(2)-mu-C(10)H(8) (M = Na, K) with 2 equiv of 1,3,5,7-cyclooctatetraene afforded a mixture of two products: M-2-COT and 2(2)-mu-COT. Compound 2(2)-mu-COT can be assembled independently in 90% yield by salt elimination upon reaction of M-2-COT with iodide 2-I-DME. The U-C(arene) distance in compound 2(2)-mu-COT is longer than that in its naphthalene counterpart K(2)-2(2)-mu-C(10)H(8)(2.822 vs 2.634 A), in accord with bonding considerations. A DFT study performed on model compounds for both M(2)-2(2)-mu-C(10)H(8) and 2(2)-mu-COT indicates that the delta bonds present in the former compound show better covalent overlap.  相似文献   
70.
The kinetics of the oxidative addition of PhSeSePh and PhTeTePh to the stable 17-electron complex *Cr(CO)3C5Me5 have been studied utilizing stopped-flow techniques. The rates of reaction are first-order in each reactant, and the enthalpy of activation decreases in going from Se (deltaH(double dagger) = 7.0 +/- 0.5 kcal/mol, deltaS(double dagger) = -22 +/- 3 eu) to Te (deltaH(double dagger) = 4.0 +/- 0.5 kcal/mol, deltaS(double dagger) = -26 +/- 3 eu). The kinetics of the oxidative addition of PhSeH and *Cr(CO)3C5Me5 show a change in mechanism in going from low (overall third-order) to high (overall second-order) temperatures. The enthalpies of the oxidative addition of PhE-EPh to *Cr(CO)3C5Me5 in toluene solution have been measured and found to be -29.6, -30.8, and -28.9 kcal/mol for S, Se, and Te, respectively. These data are combined with enthalpies of activation from kinetic studies to yield estimates for the solution-phase PhE-EPh bond strengths of 46, 41, and 33 kcal/mol for E = S, Se, and Te, respectively. The corresponding Cr-EPh bond strengths are 38, 36, and 31 kcal/mol. Two methods have been used to determine the enthalpy of hydrogenation of PhSeSePh in toluene on the basis of reactions of HSPh and HSePh with either *Cr(CO)3C5Me5 or 2-pyridine thione. These data lead to a thermochemical estimate of 72 kcal/mol for the PhSe-H bond strength in toluene solution, which is in good agreement with kinetic studies of H atom transfer from HSePh at higher temperatures. The reaction of H-Cr(CO)3C5Me5 with PhSe-SePh is accelerated by the addition of a Cr radical and occurs via a rapid radical chain reaction. In contrast, the reaction of PhTe-TePh and H-Cr(CO)3C5Me5 does not occur at any appreciable rate at room temperature, even in the presence of added Cr radicals. This is in keeping with a low PhTe-H bond strength blocking the chain and implies that H-TePh < or = 63 kcal/mol. Structural data are reported for PhSe-Cr(CO)3C5Me5 and PhS-Cr(CO)3C5Me5. The two isostructural complexes do not show signs of an increase in steric strain in terms of metal-ligand bonds or angles as the Cr-EPh bond is shortened in going from Se to S. Bond strength estimates of the PhE-H and PhE-EPh derived from density functional theory calculations are in reasonable agreement with experimental data for E = Se but not for E = Te. The nature of the singly occupied molecular orbital of the *EPh radicals is calculated to show increasing localization on the chalcogenide atom in going from S to Se to Te.  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