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   96篇
  免费   3篇
  国内免费   1篇
化学   43篇
晶体学   1篇
力学   4篇
数学   8篇
物理学   44篇
  2020年   1篇
  2019年   2篇
  2017年   2篇
  2016年   3篇
  2014年   1篇
  2013年   3篇
  2012年   2篇
  2011年   3篇
  2009年   2篇
  2007年   2篇
  2006年   1篇
  2005年   5篇
  2004年   2篇
  2003年   2篇
  2002年   2篇
  2001年   1篇
  2000年   1篇
  1999年   3篇
  1996年   11篇
  1995年   6篇
  1994年   6篇
  1993年   3篇
  1992年   7篇
  1991年   2篇
  1990年   2篇
  1989年   1篇
  1987年   5篇
  1986年   1篇
  1985年   1篇
  1981年   1篇
  1980年   1篇
  1978年   1篇
  1977年   1篇
  1975年   1篇
  1974年   2篇
  1973年   1篇
  1969年   2篇
  1968年   3篇
  1967年   2篇
  1887年   1篇
  1886年   1篇
排序方式: 共有100条查询结果,搜索用时 15 毫秒
41.
42.
The mass spectra of five selenoureas and two selenothiocarbamic esters are reported and discussed. Fragmentation modes are compared with those reported for the ureas, thioureas, carbamates and thiocarbamates.  相似文献   
43.
The formation and interconversion of the N-methylphenothiazine cation radical and the N-methylphenothiazine dication from both N-methylphenothiazine and N-methylphenothiazine S-oxide in sulfuric acid solutions have been demonstrated with e.s.r. and absorption spectroscopy. Cryoscopic measurements have shown that in slightly aqueous sulfuric acid N-methylphenothiazine S-oxide is converted to the N-methylphenothiazine dication and, analogously, phenothiazine 5-oxide is converted to the protonated phenazathionium ion (the phenothiazine dication).  相似文献   
44.
Integrated infrared cross-sections and wavenumber positions for the vibrational modes of a range of hydrofluoroethers (HFEs) and hydrofluoropolyethers (HFPEs) have been calculated. Spectra were determined using a density functional method with an empirically derived correction for the wavenumbers of band positions. Radiative efficiencies (REs) were determined using the Pinnock et al. method and were used with atmospheric lifetimes from the literature to determine global warming potentials (GWPs). For the HFEs and the majority of the molecules in the HG series HFPEs, theoretically determined absorption cross-sections and REs lie within ca. 10% of those determined using measured spectra. For the larger molecules in the HG series and the HG′ series of HFPEs, agreement is less good, with theoretical values for the integrated cross-sections being up to 35% higher than the experimental values; REs are up to 45% higher. Our method gives better results than previous theoretical approaches, because of the level of theory chosen and, for REs, because an empirical wavenumber correction derived for perfluorocarbons is effective in predicting the positions of C-F stretching frequencies at around 1250 cm−1 for the molecules considered here.  相似文献   
45.
The two-dimensional striation model consists of a nonlinear system of PDE's which arises in the modeling of the ionospheric plasma. The local-in-time existence of strong solutions is first proved using Banach's fixed point theorem. Then, under physically relevant assumptions, the system is shown to be nonlinearly unstable as soon as it is linearly unstable. Moreover, the instability occurs before the possible blow-up time of the solution. The proof relies on an earlier work of Hwang and Guo (2003 Hwang , H. J. , Guo , Y. ( 2003 ). On the dynamical Rayleigh–Taylor instability . Arch. of Rat. Mech. Anal. 167 : 235253 .[Crossref], [Web of Science ®] [Google Scholar]). The first step of the proof is to investigate under which conditions the linearized system is unstable and to prove that its spectrum is bounded, by means of a variational formulation. The second one consists in constructing a family of solutions depending on the parameter δ measuring the smallness of the perturbation to the steady-state. Thanks to the boundedness of the linearized spectrum, this family of solutions is shown to be unstable by means of a power series expansion in δ.  相似文献   
46.
Thianthrene cation radical salts, Th(*)(+) X(-)(X(-) = a, ClO(4)(-); b, PF(6)(-); c, SbF(6)(-)), add to cycloalkenes (C(5)-C(8)) in acetonitrile (MeCN) to form 1,2-bis(5-thianthreniumyl)cycloalkane salts and 1,2-(5,10-thianthreniumdiyl)cycloalkane salts, most of which have now been isolated and characterized. These are called bis- (3, 6, 9, 12) and monoadducts (4, 7, 10, 13). The proportional amount of the monoadduct obtained in the initial stage of the reaction varied with the cycloalkene in the order C(6) < C(5) < C(7) < C(8). Thus, the ratio bis:mono for C(5) and C(7) was, respectively, about 80/20 and 50/50. In contrast, only about 5% of the C(6) monoadduct (7a) and none of 7b,c was obtained, while for C(8) none of the bisadducts 12a-c was found. Bisadducts 3 and 9 lost thianthrene (Th) slowly in MeCN solution and changed into monoadducts 4 and 10. A comparable change from 6a into 7a was not observed. The monoadducts, themselves, lost a proton slowly in dry MeCN and opened into 1-(5-thianthreniumyl)cycloalkenes (5, 8, 11, 14). With 3 and 9, particularly, it was possible to follow with NMR spectroscopy the succession of changes, for example, 3 to 4 to 5. The opening of a monoadduct was made faster by adding a small amount of water to the solution. The bisadducts of 4-methylcyclohexene (15a) and 1,5-cyclooctadiene (17a) were isolated and characterized. Although a small amount of monodduct (16a) of 4-methylcyclohexene was found with NMR spectroscopy, it could not be isolated. Bis- and monoadducts were obtained also in additions of Th(*)(+) ClO(4)(-) to acyclic alkenes, in relative amounts that, again, varied with the alkene. From cis-2-butene the dominant product was the bisadduct (18), while the monoaduct (19) was characterized with NMR spectroscopy but could not be isolated. In contrast, trans-3-hexene gave mainly the monoadduct (21), while the bis adduct (20) could not be isolated. With 4-methyl-cis-2-pentene, both bis- (22) and monoadduct (23) were isolated, the former being dominant. The conversion of 18 into 19 was characterized with NMR spectroscopy. In all cycloalkene bisadducts, the configurational relationship of the two thianthrenium groups was trans, while in the monoadducts, the bonds to the single thianthrene dication were (necessarily) cis. In both bis- and monoadducts of acyclic alkenes, the configuration of the alkene was retained. The mechanisms of addition with retention of configuration, of conversion of a bis- into a monoadduct, and of opening of a monoadduct are discussed. Products were identified with a combination of NMR spectroscopy, X-ray crystallography, elemental analysis, and (for cycloalkene adducts) reaction with thiophenoxide ion.  相似文献   
47.
Under the influence of a uniform and parallel magnetic field, a ferromagnetic fiber suspended in a Newtonian fluid rotates to align with the field direction. This study examines the field-induced rotation process for an individual non-Brownian axisymmetric ellipsoid suspended in a stagnant Newtonian fluid. Theoretical predictions are derived by a perturbation analysis for the limiting case where the strength of the applied magnetic field far exceeds the saturation magnetization of the ellipsoid. Numerical calculations are performed for the more general problem of an ellipsoid with known isotropic, non-hysteretic magnetic properties, using nickel and a stainless steel as examples. The analysis encompasses materials with field-induced, nonlinear magnetic properties, distinguishing these results from the simpler cases where the particle magnetization is either independent of, or linearly dependent on, the strength of the applied external field. In this study, predictions indicate that when the ellipsoid is magnetically saturated, the particle rotation is governed by the magnetoviscous time constant, MV = s/0 M s 2 . It is found that the rotation rate depends strongly on the aspect ratio,a/b, of the ellipsoid, but only weakly on the dimensionless magnetization,M s/H 0. A geometric parameter for an ellipsoid, defined in eq. (2.5) - a, b major, minor semi-axes of an axisymmetric ellipsoid - D demagnetization tensor for an ellipsoid - D M magnetometric demagnetization tensor, the volume-average ofD P (r) - D P (r) position dependent demagnetization tensor, implicitly defined in eq. (2.12) - D xx,D yy,D zz demagnetization factors, the diagonal elements ofD. Values for ellipsoids are defined in eq. (2.15) - F (m) magnetic force exerted on a body in a magnetic field - H i ;H i magnetic field inside a ferromagnetic body; magnitude ofH i - H 0;H 0 magnetic field applied by external sources; magnitude ofH 0 - h i ;h ix,h iy Cartesian components of dimensionless internal magnetic field,h i =H i /H 0 - I moment of inertia tensor - k geometric parameter for hydrodynamic resistance of a body rotating in a Newtonian fluid given in eq. (2.3) - L (h);L z (h) hydrodynamic torque exerted on a rotating body; thez-component of the hydrodynamic torque - L (m);L z (m) magnetic torque exerted on a magnetic body in a magnetic field, eq. (2.10); thez-component of the magnetic torque - M;M the magnetization, or dipole moment density, of a magnetic material; the magnitude ofM - M s the saturation value ofM, approached by all ferromagnetic materials asH i becomes large (figure 3) - m s the dimensionless saturation magnetization,M s/H 0 - r position vector of a point within a ferromagnetic body - s dummy integration variable in eq. (2.5) - t time - U magnetoquasistatic potential energy of a magnetic body in a magnetic field, given in eq. (2.8) - u curve-fitting variable in eq. (4.1);u = logH i - V volume of a magnetic particle; for an axisymmetric ellipsoid,V = (4/3) ab 2 - x, y, z rectangular coordinate axes fixed in the ellipsoid (figure 1) - angle of inclination of the major axis of the ellipsoid with respect toH 0 - s viscosity of the Newtonian suspending medium - µ 0 the magnetic permeability of free space,µ 0 =4 · 10–7H/m - MV the magnetoviscous time constant, a characteristic time for a process involving a competition of viscous and magnetic stresses - the magnetic susceptibility of a magnetic material, = M/H i - ; z angular velocity of a rotating body; angular velocity about thez-axis of an ellipsoid, z=–d/dt  相似文献   
48.
Addition of phenoxathiin cation radical (PO*+) to acyclic alkenes in acetonitrile (MeCN) solution occurred stereospecifically to form bis(10-phenoxathiiniumyl)alkane adducts. Stereospecific trans addition is ascribed to the intermediacy of an episulfonium cation radical. The alkenes used were cis- and trans-2-butene, cis- and trans-2-pentene, cis- and trans-4-methyl-2-pentene, cis- and trans-4-octene, trans-3-hexene, trans-3-octene, trans-5-decene, cis-2-hexene, and cis-2-heptene. The erythro bisadducts (compounds 6) were obtained with trans-alkenes, while threo bisadducts (compounds 7) were obtained with cis-alkenes. The assigned structures of 6 and 7 were consistent with their NMR spectra and, in one case, 6c (the adduct of trans-4-methyl-2-pentene) was confirmed with X-ray crystallography. Additions of PO*+ to 1,4-hexa-, 1,5-hexa-, 1,6-hepta-, and 1,7-octadiene gave bis(10-phenoxathiiniumyl)alkenes (compounds 8), the assigned structures of which were consistent with their NMR spectra. Each of these adducts lost a proton and phenoxathiin (PO) when treated with basic alumina in MeCN solution. Compounds 6 (from trans-alkenes) gave mixtures of (Z)- (9) and (E)-(10-phenoxathiiniumyl)alkenes (10) in which the (Z)-isomers (9) were dominant. On the other hand, compounds 7 (from cis-alkenes) gave mixtures of 9 and 10 in which, with one exception (the adduct 7c of cis-4-methyl-2-pentene), compounds 10 were dominant. The path to elimination is discussed. The alkenes 9 and 10 were characterized with NMR spectroscopy and, in one case (9a), with X-ray crystallography. Reactions of 8b-d with basic alumina gave mixtures of (E)- (13) and (Z)-(10-phenoxathiiniumyl)dienes (14), in which compounds 13 were dominant. The configuration of the product from 8a (the adduct of 1,4-hexadiene) could not be settled. Noteworthy features in the coupling patterns and chemical shifts in the NMR spectra of some of the adducts and their products are discussed and related to adduct conformations.  相似文献   
49.
[reaction: see text] Thianthrene cation radical tetrafluoroborate (Th*+ BF4-) added to 2-butyne, 3-hexyne, 4-octyne, and 5-decyne in MeCN to form trans bisadducts R(Th+)C=C(Th+)R, where R = Me, Et, Pr, Bu (7a-d). Phenoxathiin cation radical tetrafluoroborate (PO*+ BF4-) added similarly to the last three alkynes to form adducts R(PO+)C=C(PO+)R, 8b-d. Cyclic monoadducts were not found. The trans structures of 7 and 8 were deduced with X-ray crystallography (7c) and NMR spectroscopy. When solutions of adducts in CHCl3 and MeCN were deposited on activated alumina, elimination of thianthrene (Th) and phenoxathiin (PO) occurred almost quantitatively. Detailed studies with (7b-d) indicated that a cumulene (15) was formed by the elimination of Th and that 15 was subsequently converted into small amounts of other products. In CHCl3, these products were the respective alkyne, thianthrene 5-oxide, an alpha-diketone (11), an alpha-hydroxyalkyne (12), and hydrogen. The same products were formed in MeCN along with an alpha-acetamidoalkyne (13). The formation of 15 and products derived from it is explained and was confirmed by preparation and reactions of 2,3,4-hexatriene.  相似文献   
50.
HFC-134a (CF3CH2F) is the most rapidly growing hydrofluorocarbon in terms of atmospheric abundance. It is currently used in a large number of household refrigerators and air-conditioning systems and its concentration in the atmosphere is forecast to increase substantially over the next 50-100 years. Previous estimates of its radiative forcing per unit concentration have differed significantly ∼25%. This paper uses a two-step approach to resolve this discrepancy. In the first step six independent absorption cross section datasets are analysed. We find that, for the integrated cross section in the spectral bands that contribute most to the radiative forcing, the differences between the various datasets are typically smaller than 5% and that the dependence on pressure and temperature is not significant. A “recommended' HFC-134a infrared absorption spectrum was obtained based on the average band intensities of the strongest bands. In the second step, the “recommended' HFC-134a spectrum was used in six different radiative transfer models to calculate the HFC-134a radiative forcing efficiency. The clear-sky instantaneous radiative forcing, using a single global and annual mean profile, differed by 8%, between the 6 models, and the latitudinally-resolved adjusted cloudy sky radiative forcing estimates differed by a similar amount. We calculate that the radiative forcing efficiency of HFC-134a is .  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