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   2413篇
  免费   196篇
  国内免费   363篇
化学   1649篇
晶体学   12篇
力学   215篇
综合类   62篇
数学   466篇
物理学   568篇
  2023年   32篇
  2022年   61篇
  2021年   112篇
  2020年   73篇
  2019年   45篇
  2018年   46篇
  2017年   59篇
  2016年   87篇
  2015年   83篇
  2014年   97篇
  2013年   141篇
  2012年   108篇
  2011年   147篇
  2010年   113篇
  2009年   158篇
  2008年   146篇
  2007年   156篇
  2006年   168篇
  2005年   135篇
  2004年   111篇
  2003年   117篇
  2002年   93篇
  2001年   85篇
  2000年   59篇
  1999年   69篇
  1998年   76篇
  1997年   63篇
  1996年   40篇
  1995年   34篇
  1994年   35篇
  1993年   30篇
  1992年   26篇
  1991年   12篇
  1990年   20篇
  1989年   15篇
  1988年   19篇
  1987年   12篇
  1986年   13篇
  1985年   13篇
  1984年   8篇
  1983年   2篇
  1982年   11篇
  1981年   6篇
  1980年   7篇
  1979年   7篇
  1978年   4篇
  1976年   5篇
  1973年   4篇
  1972年   4篇
  1969年   1篇
排序方式: 共有2972条查询结果,搜索用时 16 毫秒
91.
研究了低于300 ℃时两种氧化铈对稀燃阶段NOx存储性能的影响,催化剂由2%(w)Pt/Al2O3(PA)与CeO2-X(X=S,I)机械混合制备. X射线衍射(XRD),BET表面积和扫描电子显微镜(SEM)用于表征材料的物理结构. X射线光电子能谱(XPS)和H2程序升温还原(H2-TPR)用于表面Ce3+和活性氧定量. 原位漫反射傅里叶变换红外光谱(in-situ DRIFTS)用于分析表面NOx吸附物种. 相比于CeO2-I,CeO2-S 具有优良的物理化学性能,包括高比表面积、丰富的空隙结构、较高的抗老化能力及表面Ce3+浓度. 因而,Pt/Al2O3+CeO2-S 表现出优异的NOx存储能力. 此外,PA+CeO2-X(X=S,I)上存在Pt 与CeO2之间的相互作用,可提高表面氧物种的活性进而促进NO氧化及NOx存储. PA+CeO2-S上的这种相互作用要强于PA+CeO2-I. 研究表明,表面Ce3+浓度和活性氧含量对NOx存储起到重要作用. 然而经过水热处理后,Pt 与老化的氧化铈(ACS,ACI)之间的相互作用降低,并且两种氧化铈NOx存储性能显著下降. 另外,与PA+ACS(ACI)相比,PA+PACS(PACI)样品NOx存储能力得到改善,这归因于表面氧物种活性增加能促进硝酸盐的形成.  相似文献   
92.
Poly(ethylene glycol) (PEG)-supported dendrimers have been synthesized using 2.4,6-trichloro-1,3,5-triazine (TCT) as dendrons and tris(hydroxymethyl)aminomethane as tinkers with high loading capacity, excellent solubility and thermal stability by divergent method. The new synthesized PEG-supported G2.0 dendrimer has 10 times as large functional group loading capacity as commercial PEG3400 with overall yield 44.0%.  相似文献   
93.
Low-temperature heat capacities of the compound Na(C4H7O5)·H2O(s) have been measured with an automated adiabatic calorimeter. A solid-solid phase transition and dehydration occur at 290-318 K and 367-373 K, respectively. The enthalpy and entropy of the solid-solid transition are ΔtransHm = (5.75 ± 0.01) kJ mol−1 and ΔtransSm = (18.47 ± 0.02) J K−1 mol−1. The enthalpy and entropy of the dehydration are ΔdHm = (15.35 ± 0.03) kJ mol−1 and ΔdSm = (41.35 ± 0.08) J K−1 mol−1. Experimental values of heat capacities for the solids (I and II) and the solid-liquid mixture (III) have been fitted to polynomial equations.  相似文献   
94.
The aqueous reactions,
  相似文献   
95.
The molar enthalpies of the solid–solid and solid–liquid phase transitions were determined by differential scanning calorimetry for pure TbCl3 and KTb2Cl7, RbTb2Cl7, CsTb2Cl7, K3TbCl6, Rb3TbCl6 and Cs3TbCl6 compounds. Both types of compounds, i.e. M3TbCl6 and MTb2Cl7 (M=K, Rb, Cs) melt congruently and show additionally a solid–solid phase transition with a corresponding enthalpy Δtrs H 0 of 6.1, 7.6 and 7.0 kJ mol–1 for potassium, rubidium and caesium M3TbCl6 compounds andΔtrs H 0 of 17.1 (rubidium) and of 12.1 and 10.9 kJ mol–1 (caesium) for MTb2Cl7 compounds, respectively. The enthalpies of fusion were measured for all the above compounds with the exception of Rb3TbCl6 and Cs3TbCl6. The heat capacities of the solid and liquid compounds have been determined by differential scanning calorimetry (DSC) in the temperature range 300–1100 K. The experimental heat capacity strongly increases in the vicinity of a phase transition, but varies smoothly in the temperature ranges excluding these transformations. C p data were fitted by an equation, which provided a satisfactory representation up to the temperatures of C p discontinuity. The measured heat capacities were checked for consistency by calculating the enthalpy of formation of the liquid phase, which had been previously measured. The results obtained agreed satisfactorily with these experimental data. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   
96.
The heat capacities of aqueous solutions of acetone, 2,5-hexanedione, diethyl ether, 1,2-dimethoxyethane, benzyl alcohol and cyclohexanol at concentrations of 0.1 to 1.0 mol⋅kg−1 were determined at temperatures of 298.15, 423.15, 473.15 and 523.15 K and pressures up to 28 MPa. The measurements were performed at ambient conditions using the commercial Picker differential flow calorimeter and at high temperatures and pressures with a customized Picker type calorimeter constructed at the Blaise Pascal University, Clermont-Ferrand. Standard molar heat capacities were obtained by weighted extrapolation to the infinite dilution limit. The contributions of –CO–, –O– and –OH groups to the standard molar volume and standard molar heat capacity were determined from the newly determined and literature data. The variation of the three oxygen-containing group contributions with temperature and molecular structure is examined qualitatively.  相似文献   
97.
Temperature-modulated calorimetry (TMC) allows the experimental evaluation of the kinetic parameters of the glass transition from quasi-isothermal experiments. In this paper, model calculations based on experimental data are presented for the total and reversing apparent heat capacities on heating and cooling through the glass transition region as a function of heating rate and modulation frequency for the modulated differential scanning calorimeter (MDSC). Amorphous poly(ethylene terephthalate) (PET) is used as the example polymer and a simple first-order kinetics is fitted to the data. The total heat flow carries the hysteresis information (enthalpy relaxation, thermal history) and indications of changes in modulation frequency due to the glass transition. The reversing heat flow permits the assessment of the first and higher harmonics of the apparent heat capacities. The computations are carried out by numerical integrations with up to 5000 steps. Comparisons of the calculations with experiments are possible. As one moves further from equilibrium, i.e. the liquid state, cooperative kinetics must be used to match model and experiment.On leave from Toray Industries, Inc., Otsu, Shiga 520, Japan.This work was supported by the Division of Materials Research, National Science Foundation, Polymers Program, Grant # DMR 90-00520 and the Division of Materials Sciences, Office of Basic Energy Sciences, U. S. Department of Energy at Oak Ridge National Laboratory, managed by Lockheed Martin Energy Research Corp. for the U. S. Department of Energy, under contract number DE-AC05-96OR22464. Support for instrumentation came from TA Instruments, Inc. Research support was also given by ICI Paints, and Toray Industries, Inc.  相似文献   
98.
The specific heat capacities of hexamethylphosphoric triamide, diethylpropionamide, their aqueous solutions, and mixtures of hexamethylphosphoric triamide with formamide were measured in the temperature range from 288.15 to 318.15 K. The dependences of the partial molar heat capacity of aqueous solutions of amides on the composition of the mixture have maxima in the region of 0.02–0.04 molar fractions of amide. The maximum on a similar dependence for solutions of hexamethylphosphoric triamide corresponds to the concentration of 0.01 molar fractions. The conclusion on the formation of solvates (hydrates) in the systems studied was made. The heat capacity coefficients of pair and triple interactions were calculated in terms of the McMillan-Mayer theory. A change in the heat capacity characteristics with the temperature change was analyzed. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 12, pp. 2479–2483, December, 1998.  相似文献   
99.
In ternary aqueous solutions, hydrophobic solutes such as alcohols tend to aggregate with surfactants to form mixed micelles. These systems can be studied by meas of the functions of transfer of hydrophobic solutes from water to aqueous solutions of surfactant. These thermodynamic functions often go through extrema in the critical micellar concentration (CMC) region of the surfactant. A simple model based on interactions between surfactant and hydrophobic solute monomers, on the distribution of the hydrophobic solute between water and the micelles and on the shift in the CMC induced by the hydrophobic solute, can simulate the magnitude and trends of the transfer functions using parameters which are mostly derived from the binary systems. In order to check the model more quantitatively, volumes and heat capacities of transfer of alcohols from water to aqueous solutions of a nonionic surfactant, octyldimethylamine oxide, were measured. A quantitative agreement was achieved with three adjustable parameters. Good fits are also obtained for the transfers to the ionic surfactants, octylamine hydrobromide and sodium dodecylsulfate. When the equilibrium displacement contribution is small, the distribution constants and the partial molar properties of the alcohols in the micellar phase agree well with the parameters obtained with similar models.  相似文献   
100.
Equations have been derived from which the maximum sample size which may be loaded on a column without significantly affecting the performance may be calculated. It is shown that micro-packed columns can handle a sample which is as much as forty times larger than that which can be loaded onto wall-coated open tubular columns with comparable conditions.  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