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   75篇
  免费   3篇
  国内免费   5篇
化学   67篇
晶体学   1篇
物理学   15篇
  2022年   2篇
  2019年   1篇
  2018年   2篇
  2017年   3篇
  2016年   2篇
  2015年   1篇
  2014年   2篇
  2013年   5篇
  2012年   5篇
  2011年   4篇
  2010年   3篇
  2009年   5篇
  2008年   5篇
  2007年   4篇
  2006年   8篇
  2005年   4篇
  2004年   3篇
  2003年   3篇
  2002年   2篇
  2001年   2篇
  2000年   1篇
  1999年   1篇
  1998年   3篇
  1997年   2篇
  1996年   2篇
  1995年   2篇
  1994年   3篇
  1993年   1篇
  1989年   1篇
  1988年   1篇
排序方式: 共有83条查询结果,搜索用时 187 毫秒
11.
The new pyrazine-pillared solids, AgReO4(C4H4N2) (I) and Ag3Mo2O4F7(C4H4N2)3 (C4H4N2=pyrazine, pyz) (II), were synthesized by hydrothermal methods at 150 °C and characterized using single crystal X-ray diffraction (IP21/c, No. 14, Z=4, a=7.2238(6) Å, b=7.4940(7) Å, c=15.451(1) Å, β=92.296(4)°; IIP2/n, No. 13, Z=2, a=7.6465(9) Å, b=7.1888(5) Å, c=19.142(2) Å, β=100.284(8)°), thermogravimetric analysis, UV-Vis diffuse reflectance, and photoluminescence measurements. Individual Ag(pyz) chains in I are bonded to three perrhenate ReO4- tetrahedra per layer, while each layer in II contains sets of three edge-shared Ag(pyz) chains (π-π stacked) that are edge-shared to four Mo2O4F73- dimers. A relatively small interlayer spacing results from the short length of the pyrazine pillars, and which can be removed at just slightly above their preparation temperature, at >150-175 °C, to produce crystalline AgReO4 for I, and Ag2MoO4 and an unidentified product for II. Both pillared solids exhibit strong orange-yellow photoemission, at 575 nm for I and 560 nm for II, arising from electronic excitations across (charge transfer) band gaps of 2.91 and 2.76 eV in each, respectively. Their structures and properties are analyzed with respect to parent ‘organic free’ silver perrhenate and molybdate solids which manifest similar photoemissions, as well as to the calculated electronic band structures.  相似文献   
12.
Molybdate was examined as a complex-forming additive to the CE background electrolytes (BGE) to affect the selectivity of separation of polyhydric phenols such as flavonoids (apigenin, hyperoside, luteolin, quercetin and rutin) and hydroxyphenylcarboxylic acids (ferulic, caffeic, p-coumaric and chlorogenic acid). Effects of the buffer concentrations and pH and the influence of molybdate concentration on the migration times of the analytes were investigated. In contrast to borate (which is a buffering and complex-forming agent generally used in CE at pH ≥9) molybdate forms more stable complexes with aromatic o-dihydroxy compounds and hence the complex-formation effect is observed at considerably lower pH. Model mixtures of cinnamic acid, ferulic acid, caffeic acid and 3-hydroxycinnamic acid were separated with 25 mM morpholinoethanesulfonic acid of pH 5.4 (adjusted with Tris) containing 0.15 mM sodium molybdate as the BGE (25 kV, silica capillary effective length 45 cm × 0.1 mm I.D., UV-vis detection at 280 nm). With 25 mM 2-hydroxy-3-[4-(2-hydroxyethyl)-1-piperazinyl]propanesulphonic acid/Tris of pH* 7.4 containing 2 mM sodium molybdate in aqueous 25% (v/v) methanol as the BGE mixtures of all the above mentioned flavonoids, p-coumaric acid and chlorogenic acid could be separated (the same capillary as above, UV-vis detection at 263 nm). The calibration curves (analyte peak area versus concentration) were rectilinear (r > 0.998) for ≈8-35 μg/ml of an analyte (with 1-nitroso-2-naphthol as internal standard). The limit of quantification values ranged between 1.1 mg l−1 for p-coumaric acid and 2.8 mg l−1 for quercetin. The CE method was employed for the assay of flavonoids in medicinal plant extracts. The R.S.D. values ranged between 0.9 and 4.7% (n = 3) when determining luteolin (0.08%) and apigenin (0.92%) in dry Matricaria recutita flowers and rutin (1.03%) and hyperoside (0.82%) in dry Hypericum perforatum haulm. The recoveries were >96%.  相似文献   
13.
Reactions of AMoO4 and AMoO3 (A=Ca2+, Ba2+) with ammonia were investigated at 873 K<T<1123 K with the particular intention to synthesize novel oxynitride-perovskites of the general composition AMo(O,N)3 and to study their crystal structure. CaMo(O,N)3 and BaMo(O,N)3 were prepared by thermal ammonolysis of the corresponding CaMoO3 and BaMoO3 precursors at T=898 and 998 K, respectively. The structural parameters of the oxynitrides were obtained from Rietveld refinements of X-ray and neutron powder diffraction data. CaMo(O,N)3 crystallizes in the GdFeO3 distorted perovskite structure with orthorhombic space group Pbnm and a=5.5029(1) Å, b=5.5546(1) Å, c=7.8248(1) Å as determined by X-ray powder diffraction. Its O/N content refined from the neutron diffraction data corresponds to the composition CaMoO1.7(1)N1.3(1). BaMo(O,N)3 crystallizes in the cubic perovskite structure with space group Pmm and a=4.0657(1) Å as determined by X-ray powder diffraction. Transmission electron microscopy reveals a complex microstructure for both CaMoO3 and CaMoO1.7(1)N1.3(1) represented by twin domains of different orientation.  相似文献   
14.
A kinetic spectrophotometric procedure was developed for determination of submicromolar orthophosphate based on the reaction in which orthophosphate serves as a catalyst in the reduction of molybdenum, and the initial rate of molybdenum-blue formation (λmax = 780 nm) is proportional to the concentration of orthophosphate in the samples. The detection limit (3 × standard deviation of blank, n = 8) was 6 nM and the linear calibration ranged from 10 to 100 nM (r2 = 0.997). The precisions of this method were 3.3% at 10 nM and 5.4% at 50 nM (n = 8), respectively. Similar to other molybdate based methods, silica and arsenate in the samples can interfere with phosphate determination. The responses of silicate and arsenate were about 25% and 7% of that of orthophosphate, respectively, and their interferences were enhanced in the presence of phosphate in the samples due to the synergistic effect of phosphate with arsenate or silicate on the molybdate reagent.  相似文献   
15.
The influence of fluoride on the luminescence of LiEuM2O8 (M=Mo, W) was studied. LiEuMo2O8 and LiEuW2O8 formed the whole range of solid solutions, which emitted intense red luminescence under the excitations by 395, 465 and 535 nm wavelengths. When doped with fluoride, the materials also formed solid solutions and the luminescent intensity was remarkably enhanced. The phosphor with optimized compositions in this system would be a promising red component for solid-state lighting devices based on GaN light-emitting diodes.  相似文献   
16.
Anodic layer growth on 2024 aluminium alloy at 70 °C, under 40 V, during 60 min, in 50 g L−1 di-sodium tetraborate solution containing di-sodium molybdate from 0.1 to 0.5 M (pH 10) is examined. Anodising behaviours strongly depend on additive concentration. Development of anodic films is favoured with weak molybdate additions (<0.3-0.4 M). The film thicknesses increase and the porosity of anodic layers decreases. Molybdenum (+VI), detected by X-ray photoelectron spectroscopy (XPS) analysis, is present in the anodic films and the Mo incorporation, studied by energy dispersive spectroscopy (EDS) analysis, increases with molybdate concentration. However, for high molybdate concentrations (>0.4 M), anodising behaviour becomes complex with the formation of a blue molybdenum oxide at the cathode. The growth of aluminium oxide is hindered. As the anodic layers are thinner, the Mo(+VI) incorporation significantly decreases. These two configurations implicate different corrosion performances in 5% sodium chloride solution at 35 °C. As the alkaline anodic layer formed with 0.3 M molybdate species is the thickest and the Mo incorporation is the more pronounced, its corrosion resistance is the highest. The effect of morphology and composition of anodic films on pitting corrosion is also discussed.  相似文献   
17.
The kinetic booster effect of dimethylsulfoxide on the chemical generation of singlet oxygen, 1O2, from the disproportionation of hydrogen peroxide catalyzed by molybdate ions in methanol has been evidenced by detection of the IR luminescence of 1O2 at 1270 nm and by 95Mo NMR spectroscopy. DMSO interacts rapidly, through a direct oxygen transfer with the stable tetraperoxomolybdate , leading to DMSO2 and to the unstable triperoxomolybdate , which releases 1O2. The procedure was applied to accelerate the dark singlet oxygenation of β-citronellol and α-terpinene.  相似文献   
18.
(Me2NH2)[(Ph3Sn)3(MoO4)2], a Triorganotin Molybdate with Layer Structure The reaction of [(Ph3Sn)2MoO4] with (Me2NH2)Cl in an acetonitrile/water mixture leads to the formation of (Me2NH2)[(Ph3Sn)3(MoO4)2] ( 1 ). ( 1 ) crystallizes in the space group Pca21 with a = 1967.0(4), b = 1353.1(2) and c = 2176.6(5) pm. In the crystal structure of 1 Ph3SnO2 bipyramides and MoO4 tetrahedra are linked by corner sharing to give a layer structure. Additionally the layers are connected by O···H···N hydrogen bridges between MoO4 groups and [Me2NH2]+ ions to give a 3D network structure.  相似文献   
19.
铑(Ⅲ)-钼酸盐-罗丹明B显色体系的研究和应用   总被引:1,自引:0,他引:1  
在聚乙烯醇(PVA)存在下,铑(Ⅲ)与钼酸盐和罗丹明B(RB)形成离子缔合物,缔合物的最大吸收位于580nm,摩尔吸光系数ε值1.68×105L·mol-1·cm-1,服从比耳定律范围0~10μgRh/25mL,方法用于催化剂和冶金产品中铑的测定,结果满意  相似文献   
20.
Frank C  Schroeder F  Ebinghaus R  Ruck W 《Talanta》2006,70(3):513-517
A sequential injection analysis system (SIA) is described which is suited for the fast determination of filterable molybdate reactive phosphate (FRP, 0.2 μm) in coastal waters. It processes up to 270 samples per hour with a detection limit (3σ) of 0.05 μM and is used for surface mapping of phosphate in areas with steep concentration gradients like the Wadden Sea. The determination is based on the reaction of phosphate with acidic molybdate to phosphomolybdate, which builds non-fluorescent ion pairs with rhodamine 6G. The remaining rhodamine fluorescence is detected at 550 nm with an excitation at 470 nm. Syringe pump, valve and detector were controlled by a self made python programme, which was optimised for high speed SIA measurements in monitoring applications.  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