首页 | 官方网站   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   33258篇
  免费   5806篇
  国内免费   3920篇
数理化   42984篇
  2024年   82篇
  2023年   696篇
  2022年   836篇
  2021年   1138篇
  2020年   1377篇
  2019年   1301篇
  2018年   1196篇
  2017年   1141篇
  2016年   1597篇
  2015年   1588篇
  2014年   1880篇
  2013年   2446篇
  2012年   2974篇
  2011年   3140篇
  2010年   2126篇
  2009年   1987篇
  2008年   2232篇
  2007年   1926篇
  2006年   1735篇
  2005年   1485篇
  2004年   1132篇
  2003年   958篇
  2002年   955篇
  2001年   762篇
  2000年   654篇
  1999年   758篇
  1998年   611篇
  1997年   588篇
  1996年   546篇
  1995年   526篇
  1994年   434篇
  1993年   375篇
  1992年   318篇
  1991年   279篇
  1990年   273篇
  1989年   205篇
  1988年   157篇
  1987年   134篇
  1986年   98篇
  1985年   108篇
  1984年   69篇
  1983年   56篇
  1982年   54篇
  1981年   25篇
  1980年   12篇
  1979年   3篇
  1976年   1篇
  1959年   1篇
  1957年   8篇
  1936年   1篇
排序方式: 共有10000条查询结果,搜索用时 31 毫秒
991.
Thin film composite (TFC) membranes based polyamide were prepared with m-phenylenediamine (MPD), m-phenylenediamine-5-sulfonic acid (SMPD) and trimesoyl chloride (TMC) through interfacial polymerization technique on the polysulphone supporting film. The membranes were characterized using permeation experiments with salt water, attenuated total reflectance infrared (ATR-IR) and X-ray photoelectronic spectroscopy (XPS) as well as scanning electronic microscopy (SEM). This study has shown that the active layer of TFC membrane is aromatic polyamide, including sulfuric acid function group (-SO3H) according to the result of ATR-IR and XPS. The NaCl rejection of RO membranes decreased and the flux increased when WSMPD/WMPD increased from 0 to 1, and the linear part with pendant -COOH in membrane barrier layer increased with the increase of SMPD content, but the surface of membrane becoming smoother and smoother with the increase of SMPD content. So the membranes performance mainly was determined by chemical structure in their barrier layer.  相似文献   
992.
用光电化学方法研究棒状和多孔氧化银电极在阳极极化过程中的光响应规律可以得到许多信息。最大光响应出现在电极表面的AgO被充分地还原为Ag_2O以及Ag结晶即将生成瞬间, 多孔电极和实体电极开路光电位ΔV_(ph,oc)之比有助于对多孔电极孔结构的了解, 首次观测到n-p-n光响应波形的转化。  相似文献   
993.
单分散二氧化钛超微粒子的制备   总被引:10,自引:0,他引:10  
以四丁氧基钛为原料,采用溶胶-凝胶法制备了超微二氧化钛粉末.改变热处理气氛、升温速率、水与四丁氧基钛的摩尔比以及溶剂,分别得到7nm球形单相锐钛矿以及四方形(40nm×10nm)、球形(44nm)的主相金红石超微粒子.  相似文献   
994.
[reaction: see text] We present an ab initio study of the acid-promoted hydrolysis reaction mechanism of N-formylaziridine in comparison with formamide. Since the rate of amide hydrolysis reactions depends on the formation of the tetrahedral intermediate, we focused our attention mainly on the reactant complex, the tetrahedral intermediate, and the transition state connecting these two stationary points. Geometries were optimized using the density functional theory, and the energetics were refined using ab initio theory including electron correlation. Solvent effects were investigated by using polarizable continuum method calculations. The proton-transfer reaction between the O-protonated and N-protonated amides was investigated. In acidic media, despite that the N-protonated species is more stable than the O-protonated one, it is predicted that both N-protonated and O-protonated pathways compete in the hydrolysis reaction of N-formylaziridine.  相似文献   
995.
Separation of the dyes methyl violet, methylene blue, and congo red from aqueous solutions by paper capillary permeation adsorption method was studied using paper. Nearly 100% of the investigated dyes could be separated under the optimum conditions. The effect of pH on the separation efficiency was studied in particular. At pH 5–9, 1.3–11, and 7–11, the maximum separation was achieved for methyl violet, methylene blue and congo red, respectively. The effects of dye concentration and some foreign ions on the separatability were examined. Moreover, the selective separation of some dyes was attempted by elution with chemical reagents.  相似文献   
996.
Low-valent aluminum Al(i) chemistry has attracted extensive research interest due to its unique chemical and catalytic properties but is limited by its low stability. Herein, a hourglass phosphomolybdate cluster with a metal-center sandwiched by two benzene-like planar subunits and large steric-hindrance is used as a scaffold to stabilize low-valent Al(i) species. Two hybrid structures, (H3O)2(H2bpe)11[AlIII(H2O)2]3{[AlI(P4MoV6O31H6)2]3·7H2O (abbr. Al6{P4Mo6}6) and (H3O)3(H2bpe)3[AlI(P4MoV6O31H7)2]·3.5H2O (abbr. Al{P4Mo6}2) (bpe = trans-1,2-di-(4-pyridyl)-ethylene) were successfully synthesized with Al(i)-sandwiched polyoxoanionic clusters as the first inorganic-ferrocene analogues of a monovalent group 13 element with dual Lewis and Brønsted acid sites. As dual-acid catalysts, these hourglass structures efficiently catalyze a solvent-free four-component domino reaction to synthesize 1,5-benzodiazepines. This work provides a new strategy to stabilize low-valent Al(i) species using a polyoxometalate scaffold.

Monovalent aluminum(i) species was successfully stabilized using a reduced phosphomolybdate scaffold as a dual-acid catalyst for a four-component domino reaction.

Low- or sub-valence aluminum compounds are increasingly growing into a significant frontier subject in coordination and modern organic synthetic chemistry owing to their unique singlet carbene character, Lewis acid/base properties and catalytic reactivity.1 However, low-valence aluminum(i) compounds have inherent electron deficiency and exhibit thermodynamic instability, making them prone to self-polymerization with metal–metal bonds2 or disproportionation3 to metallic Al and Al(iii) species. Inspired by the special stabilizing effect of metallocene compounds, a ligand stabilization strategy has recently been undertaken to stabilize the low-valence aluminum center.4,5 In this regard, the utilized ligand should satisfy two key criteria: (i) sufficient steric hindrance is required to inhibit monomer polymerization; and (ii) a suitable electronic effect is needed to stabilize the aluminum(i) center. A few organometallic Al(i) compounds protected by bulky organic groups have been prepared such as [(Cp*Al)4] (Cp* = C5Me5),6 and [(CMe3)3SiAl4].7 However, despite having the ligand effect, most of these Al(i) compounds still decompose in aqueous solutions or heating conditions. In contrast to organometallic Al(i) compounds, inorganic Al(i) structures, i.e. monomeric monohalides, only exist in gaseous form at high temperature8 and to the best of our knowledge, no stable inorganic Al(i) compound is known at room temperature due to thermodynamic instability. Therefore, exploring efficient strategies to synthesize stable inorganic Al(i) compounds remains highly desired but a great challenge.Polyoxometalates (POMs), a diverse family of inorganic molecular clusters based on early-transition metals (W, Mo, V, Nb, and Ta), have extensively attracted attention in research in various fields of materials science, coordination chemistry, medicinal chemistry and catalysis science.9–11 Owing to their adjustable constituent elements and well-defined structures, POMs have been considered as promising inorganic ligands to stabilize high- and low-valent metal ions. For instance, Rompel et al.12 reported one Keggin-type [α-CrW12O40]5− anion in which a labile {CrIIIO4} tetrahedral unit was assembled at the center of the cluster. Li and co-workers employed a monolacunary Keggin-type inorganic ligand to stabilize a high-valent Cu3+ ion.13 As a unique member of the POM family, the hourglass-type phosphomolybdate cluster {M[P4MoV6O31]2}n (abbr. M{P4Mo6}2), consisting of two [P4MoV6O31]12− (abbr. {P4Mo6}2) subunits bridged by one metal (M) center, represents a fully reduced metal-oxo cluster. With all Mo atoms in the oxidation state of (+5), a more negative charge is endowed to the cluster surface.14,15 Such high electron density of {M[P4MoV6O31]2}n polyoxoanions provides an electron-rich local environment for the possible stabilization of unusual-valence metals. It is worth noting that the [P4MoV6O31]12− subunit presents near-planar triangular structures with the side sizes ranging from 7.50–7.92 Å (Fig. S1). The structural feature can supply sufficient steric hindrance to restrain the polymerization of low-valence metal species. Moreover, the six Mo atoms in each [P4MoV6O31]12− subunit arrange in a planar hexagonal-ring structure like a benzene ring, implying that such {M[P4MoV6O31]2}n clusters may have a similar delocalized electron structure to conjugated benzene or cyclopentadiene. These features make [P4MoV6O31]12− a promising candidate with respect to organic protecting groups to construct an inorganic ‘ferrocene’ analogue of Al(i) (Scheme 1). Therefore, we hypothesize that hourglass-type polyoxoanion clusters are promising to stabilize the labile Al(i) center and isolate inorganic Al(i) species.Open in a separate windowScheme 1Similar ferrocene-like sandwich structure features of an inorganic hourglass-type [AlI(P4MoV6O31)2]23− polyanion to an organometallic [(η5-Cp*)2AlI]+ cation.Herein, we show a [P4MoV6O31]12− cluster as an inorganic scaffold to stabilize the Al(i) center in two hybrid compounds, (H3O)2(H2bpe)11[AlIII(H2O)2]3{[AlI(P4MoV6O31H6)2]3·7H2O (abbr. Al6{P4Mo6}6) and (H3O)3(H2bpe)3[AlI(P4MoV6O31H7)2]·3.5H2O (abbr. Al{P4Mo6}2) (bpe = trans-1,2-di-(4-pyridyl)-ethylene), in which the labile Al(i) center is sandwiched by two [P4MoV6O31]12− sides, forming an inorganic moiety of a ‘ferrocene’ analogue. Both Al6{P4Mo6}6 and Al{P4Mo6}2 are experimentally determined at room temperature for the first time, and prepared by hydrothermal reactions of Na2MoO4·2H2O, H3PO4, AlCl3·6H2O, ethanol and N-containing bpe at 160 °C with slightly different pH values. Notably, the combination of ethanol, N-containing bpe and high hydrothermal temperature is a prerequisite to the isolation of Al(i) species. First, both ethanol and N-containing bpe were used to provide a reducing environment under hydrothermal conditions. By combining high temperature and pressure, sufficient energy is supplied to reduce Mo6+ and Al3+ ions to Mo5+ and Al+ species, respectively. Then, Mo5+ species and phosphoric acid molecules are assembled to form [P4Mo6O31]12− subunits, which are subsequently combined with Al+ ions to form hourglass-type [Al(P4Mo6O31)2]23−, hence effectively stabilizing Al(i) species (Fig. 1). From the perspective of stereochemistry, two highly negative [P4Mo6O31]12− fragments, resembling the methyl cyclopentadiene organic group, sandwich one low-valent metal Al(i) center. Hence, the construction of a strong reducing hourglass-like skeleton makes it possible to stabilize the existing Al+ species.Open in a separate windowFig. 1Ball-and-stick diagram showing the assembly of the hourglass-type cluster {Al(P4Mo6)2}.Single crystal X-ray diffraction revealed the hourglass-type {Al(P4Mo6)2} cluster in Al6{P4Mo6}6 and Al{P4Mo6}2 (Table S1), in which the [P4Mo6O31]12− subunits have a C3 symmetry and display a near-planar structure formed by six edge-sharing {MoO6} octahedra with alternating short Mo–Mo single bonds and long non-bonding Mo⋯Mo contacts. The side sizes of the {P4Mo6} subunit range from 7.50–7.92 Å, which supplies sufficient steric hindrance to restrain the polymerization or disproportionation of low-valence Al(i) species. All Mo atoms are in a reduced oxidation state of +5 and the central Al atoms are in the +1 oxidation state, as confirmed by bond valence calculations (Table S2). Thus, the synthesized Al{P4Mo6}2 represents a fully reduced metal–oxygen cluster. Moreover, the six Mo atoms in each {P4Mo6} subunit present a benzene-like planar hexagonal-ring structure with a similar π-type delocalization electron interaction with Al(i) instead of organic bulky groups. Such π-type delocalization electron interaction constructs an inorganic ‘ferrocene’ analogue of Al(i) and produces sufficient delocalization energy to stabilize Al(i) species. Considering the formation mechanism of traditional metallocenes, {P4Mo6} subunits with a similar strong electron-donating ability and suitable steric-hindrance effect on Cp rings, augment the stability of Al(i) species. Al6{P4Mo6}6 and Al{P4Mo6}2 compounds also present the first isolation of aluminum-sandwiched hourglass-type clusters in POM chemistry. Importantly, regarding the inherent and strong hydrolysis of aluminum species in water, these low-valent Al(i)-containing clusters represent the first example of stable solid-state inorganic sub-valent Al(i) compounds at room temperature.The asymmetric structure of Al6{P4Mo6}6 consists of two crystallographically independent {Al(P4Mo6)2} clusters sandwiched by central Al(1) and Al(4) atoms, two bridging [Al(H2O)2]3+ (Al(2) and Al(3)) cations and six protonated bpe cations (Fig. S2). Aluminum centers involve two kinds of oxidation states: the central Al(1) and Al(4) are in the +1 state, while the bridging Al(2) and Al(3) are in the +3 state. Both Al(1) and Al(4) display the six-coordinated octahedral configuration and bridge two {P4Mo6} subunits to form two {AlI(P4Mo6)2} clusters. The average lengths of Al–O bonds are 2.318–2.324 Å for Al(1) and Al(4) (Table S3), which are slightly longer than those of classic Al–O bonds (1.90 Å) for Al(2) and Al(3), but close to that of the Al–O bond in silica-supported alkylaluminum(i) composites.16–20 The long Al–O lengths for Al(1) and Al(4) centers may be ascribed to the lower electron cloud density located at the surface of the Al(i) cation, resulting in slightly longer bonds with the surrounding oxygen donors.5,21 Moreover, the small distorted extents (sum((dijdave)/dave)2/coordination number) of {Al(1)O6} (3.86 × 10−4) and {Al(4)O6} (1.89 × 10−3) indicate that they are in regular octahedral geometry. Moreover, another structural feature of Al6{P4Mo6}6 is that {AlI(P4Mo6)2} clusters are connected by bridging [Al(H2O)2]3+ cationic fragments (Al(2) and Al(3)), forming an unusual chain-like arrangement (Fig. 2a). It is worth noting that the 1-D chain contains a large repeating monomer with the maximum spacing of 81.69 Å, consisting of twelve Al-containing fragments ({–Al2–Al1–Al3–Al4–Al3–Al1–Al2–Al1–Al3–Al4–Al3–Al1–}). Such a long repeating monomer is rare. Each repeating monomer has two types of symmetric systems: Al(2) in the middle of the monomer plays a center of mirror symmetry and divides the whole repeating monomer into two equidistant half-units of {–Al1–Al3–Al4–Al3–Al1–}; Al(4) in each half-unit further acts as the reverse symmetric center of two {–Al3–Al1–Al2–} subunits. The two types of symmetrical systems form the infinitely extending chain-like structure in Al6{P4Mo6}6. Since bpe is a rigid and conjugated molecular structure, an effective π⋯π stacking interaction emerges and results in a honeycomb-like supramolecular organic moiety, which accommodates these 1-D inorganic chains and stabilizes the whole Al6{P4Mo6}6 framework (Fig. S3 and S4).Open in a separate windowFig. 2(a) One-dimensional (1D) inorganic structure in Al6{P4Mo6}6 with a length of repeating units of 81.69 Å, consisting of twelve Al-containing fragments ({–Al2–Al1–Al3–Al4–Al3–Al1–Al2–Al1–Al3–Al4–Al3–Al1–}). (b) Four kinds of coordination environments of {AlO6} octahedra, respectively (i = 1 − x, y, 0.5 − z; ii = 0.5 − x, 1.5 − y, 1 − z).Al{P4Mo6}2 has a similar structure to Al6{P4Mo6}6 (Table S4), wherein the most obvious difference is that {AlI[P4Mo6]2} clusters exist in isolated form and interact with the surrounding protonated bpe cations via hydrogen bonding to form into a 3-D supramolecular framework (Fig. S5 and S6). The different peripheral environment around the {AlI[P4Mo6]2} cluster can affect its acidity and catalytic activity.The solid-state 27Al NMR spectrum of Al6{P4Mo6}6 depicts two distinct resonances at δ = −22.34 and 27.33 ppm due to the octahedrally coordinated AlIII and AlI sites, respectively (Fig. 3a), indicating two types of Al local environments in Al6{P4Mo6}6. In contrast, Al{P4Mo6}2 displays only one sharp signal at δ = 7.20 ppm due to the octahedrally coordinated AlI sites (Fig. 3b). The observed narrow peak-width corresponds to the highly symmetric charge distribution at the aluminum nucleus, similar to the ferrocene analogue [(η5-Cp*)2AlI]+.5 Noticeably, AlI resonance in Al6{P4Mo6}6 appears at a lower magnetic field compared to Al{P4Mo6}2, due to the different peripheral environment around the hourglass {Al(P4Mo6)2} cluster. XPS spectra of Al6{P4Mo6}6 and Al{P4Mo6}2 further affirm the valence states of Al and Mo elements (Fig. S7 and Table S5). The Al 2p XPS profile of Al6{P4Mo6}6 reveals two peaks at 74.39 and 73.75 eV ascribed to AlIII and AlI, respectively (Fig. 3c). The area ratio of the two peaks is close to 1 : 1, in consistence with the chemical structure of Al6{P4Mo6}6. The high-resolution Al 2p XPS spectrum of Al{P4Mo6}2 displays a weaker broad peak attributed to the low amount of Al+ (Fig. 3d). Moreover, the structural stabilities of Al6{P4Mo6}6 and Al{P4Mo6}2 were investigated by soaking them in water for 24 hours. Fig. S9–S11 show the comparison of XRD, IR and XPS spectra of Al6{P4Mo6}6 and Al{P4Mo6}2 before and after soaking in water. It can be found that the characteristic diffraction peaks in XRD after soaking for 24 hours still show good agreement with the simulated data (Fig. S9). The characterized absorption bands in IR spectra also exhibit good match with the original Al6{P4Mo6}6 and Al{P4Mo6}2 (Fig. S10). The XPS spectra of Al6{P4Mo6}6 after soaking in water were also obtained. There is basically no change in the high-resolution spectra of Al 2p with the AlI/AlIII atomic ratios of ca. 1 : 1 (Fig. S11). The spectroscopic and theoretical observations verify that the low valence Al(i) species can stably exist in the reduced phosphomolybdates in the solid state (Fig. S12 and Table S6). Moreover, the acidities of Al6{P4Mo6}6 and Al{P4Mo6}2 were measured to be 0.27 and 0.442 mmol g−1, respectively, demonstrating the promising potential of Al6{P4Mo6}6 and Al{P4Mo6}2 as dual-acid catalysts.Open in a separate windowFig. 3(a and b) 27Al NMR spectra of solid Al6{P4Mo6}6 and Al{P4Mo6}2; (c and d) XPS spectra of Al in Al6{P4Mo6}6 and Al{P4Mo6}2.The catalytic performance of Al6{P4Mo6}6 and Al{P4Mo6}2 was evaluated via a solvent-free four-component domino reaction for the synthesis of pharmaceutical intermediate 1,5-benzodiazepine (Table 1). With Al6{P4Mo6}6 and Al{P4Mo6}2 as catalysts, the yields of the final product 8aaa reach 83% and 75%, respectively (Table 1, entries 1 and 2). Almost no 8aaa is observed without the acid catalysts, even when the reaction is set for a long time (Table 1, entry 3). This clarifies the excellent catalytic performance of Al6{P4Mo6}6 and Al{P4Mo6}2. Typical Brønsted acid p-TsOH and Lewis acid AlCl3 as control samples yield only 43% and 29% 8aaa, respectively (Table 1, entries 4 and 5), much lower than those attained by Al6{P4Mo6}6 and Al{P4Mo6}2 catalysts. Moreover, (H2en)12[{Na0.8K0.2(H2O)}2{Na[Mo6O12(OH)3(HPO4)2(PO4)2]2}2]·7H2O22,23 (abbr. {Na[P4Mo6]2}) in contrast achieved 72% yield of 8aaa in 30 min, slower than that of Al6{P4Mo6}6 and Al{P4Mo6}2. This indicates the advantage of the unique dual-acid features of Al(i)-stabilized reduced phosphomolybdate clusters with multiple Lewis and Brønsted acid active centers, in which the synergistic effect between the Al species and reduced phosphomolybdate cluster contributes to the catalytic activity.Comparison tests of one-pot synthesis of 1,5-benzodiazepine 8aaavia a four-component domino reactiona
EntryCatalystb t 1 f (h)Yieldc (%) 3a t 2 f (h)Yieldd (%) 5aa T 3 (°C) t 3 f (min)Yielde (%) 8aaa
1 Al6{P4Mo6}6 3.0 98 1.8 92 25 20 83
2Al{P4Mo6}23.2972.089252075
3No catalyst7.0985.56225120Trace
4 p-TsOH4.0923.073252643
5AlCl34.5943.082255829
6{Na[P4Mo6]2}3.5952.586253072
Open in a separate windowaOne-pot reaction conditions: acetophenone 1a (1.00 mmol), N,N-dimethylformamide dimethyl acetal 2 (1.00 mmol), 1,2-phenylenediamine 4a (1.00 mmol), ethyl pyruvate 6a (1.00 mmol) and catalyst (10.00 mg) for the four-component domino reaction.bCatalyst (10.00 mg).cIsolated yield in the first step.dTotal isolated yield for the first two steps.eOverall isolated yield for the 3 steps.fThe time taken for the reaction to complete.Furthermore, the Al6{P4Mo6}6 catalyst displays a wide substrate scope of auto-tandem catalytic reactions. A series of functional groups including carboxyl, ester and acyl groups on the 2-position of the seven-membered rings can be smoothly converted into the desired 1,5-benzodiazepine products with high and even excellent yields (Table S7). 1,2-Phenylenediamines 4 which contain both electron-deficient (p-Cl and p-Br) and electron-rich (p-Me and 3,4-di(Me)) 1,2-phenylenediamines also undergo the reaction smoothly, providing the corresponding products in high yields within the given reaction times (Table S7).Additionally, the Al6{P4Mo6}6 catalyst can be easily recovered by simple filtration. No significant decay in the catalytic activity or selectivity was observed even after 5 recycles of Al6{P4Mo6}6 (Fig. S14). The acquired XRD pattern, and IR and XPS spectra after 5 runs further revealed the good structural integrity and high solid-state stability of Al6{P4Mo6}6 (Fig. S15–S17). Accordingly, the Al6{P4Mo6}6 cluster coupled with dual-acid sites presents great potential application towards the four-component domino reaction.In summary, two cases of low valence Al-centered hourglass-type phosphomolybdates have been reported for the first time. {P4Mo6} subunits with highly negative charge and a benzene-like planar hexagonal-ring structure, display a similar π-type electron interaction with Al(i) to construct inorganic ‘ferrocene’ analogues of Al(i), thus effectively stabilizing Al(i) species. Al(i)-POM structures are confirmed and characterized using 27Al NMR and XPS spectra. When used as acid catalysts, both Al6{P4Mo6}6 and Al{P4Mo6}2 efficiently catalyze a solvent-free domino reaction to synthesize 1,5-benzodiazepines with high yield and selectivity. The Al(i)-stabilized reduced POM structures also exhibit excellent substrate compatibility and cycle stability. The design, synthesis and successful stabilization of the subvalent metallic aluminum compounds in the solid state unravel the significance of this study. This work is also important to develop highly active and multifunctional catalysts for organic reactions.  相似文献   
997.
我们用精密自动绝热量热计测定了几种不同吸附水含量的水/硅胶吸附体系在200~320 K温度范围内的热容. 结果表明, 当吸附水含量使表面复盖度(θ)大于1时, 在相应的C_p~T曲线上会出现吸附水的相变峰. 这说明吸附在硅胶表面上的水分子已经形成了聚集态; 而当θ<1时, 由于尚未形成聚集态水, 故没有相变过程出现, 其C_p~T曲线呈光滑状. 这些现象与H_2O/γ-Al_2O_3吸附体系是一致的. 又由于硅胶表面对水分子的吸附力较γ-Al_2O_3的要小, 故在同样的吸附量的C_p~T曲线上, 水/硅胶的峰要比H_2O/γ-Al_2O_3的尖锐, 且蜂温增高的速度要快. 这些都表明, 吸附在硅胶表面上的二维表相水会随吸附量的增加而以较快的趋势接近于体相水. 此外, 由不同含水量的C_p~T曲线外推, 求出了不含吸附水的硅胶在200~300 K范围内的热容.  相似文献   
998.
The Cs‐symmetric fullerene chlorohydrin C60(Cl)(OH)(OOtBu)4 reacts with 4‐dimethylaminopyridine (DMAP) and 1,4‐diazabicyclo[2.2.2]octane (DABCO) to yield two isomers with the formula C60(O)(OOtBu)4 in good yields. These isomers differ with respect to the location of the epoxy functionality. The one from DMAP is Cs symmetric, whereas that from DABCO is C1 symmetric with the epoxy group on the central pentagon. Two different mechanisms are proposed to explain the chemoselectivity of these reactions. The reaction with DMAP involves single‐electron transfer as the key step; DMAP acts as the electron donor. A combination of an oxygen‐atom shift and SN2′′ processes (boomerang substitution) are responsible for the formation of isomer with DACBO. Various related reactions support the proposed mechanisms. The structures of new fullerene derivatives were determined by spectroscopy, single‐crystal X‐ray analysis, and chemical correlation experiments.  相似文献   
999.
DNA microarray: a high throughput approach for methylation detection   总被引:7,自引:0,他引:7  
We described a DNA microarray-based method combined with bisulphite treatment of DNA and regular PCR to examine hyper-methylation in promoter 1A of APC gene. A set of oligonucleotide probes were designed and immobilized on the aldehyde-coated glass slides for detecting the methylation pattern of 15 selected CpG sites in the region. The methylation status of 30 colorectal tumor samples have been examined by both of methylation-specific PCR (MS-PCR) and the present microarray method. The methylation pattern of the 15 CpG sites for the samples have been obtained with the microarray. A total of 19 samples out of 30 were methylated by microarray, in which five samples cannot be detected by MS-PCR due to the methylated CpG patterns not accordant to the MS-PCR primers. The detecting ratio for methylation of APC gene of colorectal tumor samples increased from 46.7% with MS-PCR to 63.3% with the microarray, which successfully demonstrated that DNA microarray-based method not only can obtained the methylation patterns for the related genes, but also decrease the false-negative results of methylation status by the conventional MS-PCR for the investigated genes.  相似文献   
1000.
Zhang J  Zhou X  Cai R  Weng L 《Inorganic chemistry》2005,44(3):716-722
The direct reactions of (C5H5)2LnCl with LiN=C(NMe2)2 proceeded at room temperature in THF under pure nitrogen to yield the lanthanocene guanidinate complexes [(C5H5)2Ln(mu-eta1:eta2-N=C(NMe2)2)]2 (Ln = Gd (1), Er (2)). Treatment of phenyl isocyanate with complexes 1 and 2 results in monoinsertion of phenyl isocyanate into the Ln-N(mu-Gua) bond to yield the corresponding insertion products [(C5H5)2Ln(mu-eta1:eta2-OC(N=C(NMe2)2)NPh)]2 (Ln = Gd (3), Er (4)), presenting the first example of unsaturated organic small molecule insertion into the metal-guanidinate ligand bond. Further investigations indicate that N,N'-diisopropylcarbodiimide does not react with complexes 1 and 2 under the same conditions; however, it readily inserts into the lithium-guanidinate ligand bond of LiN=C(NMe2)2. As a synthon of the insertion product Li[(iPrN)2C(N=C(NMe2)2)], its reaction with (C5H5)2LnCl gives the novel organolanthanide complexes containing the guanidinoacetamidinate ligand, (C5H5)2Ln[(iPrN)2C(N=C(NMe2)2)] (Ln = Yb (5), Er (6), Dy (7)). All complexes were characterized by elemental analysis and spectroscopic properties. The structures of complexes 1, 3, 5 and 7 were determined through X-ray single-crystal diffraction analysis.  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-23

京公网安备 11010802026262号