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   6725篇
  免费   504篇
  国内免费   404篇
化学   4184篇
晶体学   75篇
力学   194篇
综合类   22篇
数学   543篇
物理学   1297篇
综合类   1318篇
  2024年   10篇
  2023年   60篇
  2022年   132篇
  2021年   173篇
  2020年   182篇
  2019年   141篇
  2018年   133篇
  2017年   147篇
  2016年   251篇
  2015年   237篇
  2014年   319篇
  2013年   433篇
  2012年   548篇
  2011年   531篇
  2010年   398篇
  2009年   367篇
  2008年   455篇
  2007年   423篇
  2006年   365篇
  2005年   310篇
  2004年   306篇
  2003年   284篇
  2002年   363篇
  2001年   206篇
  2000年   170篇
  1999年   114篇
  1998年   52篇
  1997年   47篇
  1996年   48篇
  1995年   31篇
  1994年   30篇
  1993年   41篇
  1992年   25篇
  1991年   25篇
  1990年   23篇
  1989年   34篇
  1988年   20篇
  1987年   19篇
  1986年   11篇
  1985年   9篇
  1984年   12篇
  1983年   9篇
  1982年   9篇
  1981年   13篇
  1980年   16篇
  1979年   18篇
  1978年   20篇
  1977年   12篇
  1976年   9篇
  1975年   8篇
排序方式: 共有7633条查询结果,搜索用时 15 毫秒
91.
Excitation-energy dependence of fluorescence intensity and fluorescence lifetime has been measured for 4-dimethylaminobenzonitrile (DMABN), 4-aminobenzonitrile (ABN), 4-diisopropylaminobenzonitrile (DIABN), and 1-naphthonitrile (NN) in a supersonic free jet. In all cases, the fluorescence yield decreases rather dramatically, whereas the fluorescence lifetime decreases only moderately for S1 (pi pi*, L(b)) excess vibrational energy exceeding about 1000 cm(-1). This is confirmed by comparison of the normalized fluorescence excitation spectrum with the absorption spectrum of the compound in the vapor phase. The result indicates that the strong decrease in the relative fluorescence yield at higher energies is due mostly to a decrease in the radiative decay rate of the emitting state. Comparison of the experimental results with the TDDFT potential energy curves for excited states strongly suggests that the decrease in the radiative decay rate of the aminobenzonitriles at higher energies is due to the crossing of the pi pi* singlet state by the lower-lying pi sigma*(C[triple bond]N) singlet state of very small radiative decay rate. The threshold energy for the fluorescence "break-off" is in good agreement with the computed energy barrier for the pi pi*/pi sigma* crossing. For NN, on the other hand, the observed decrease is in fluorescence yield at higher excitation energies can best be attributed to the crossing of the pi pi* singlet state by the pi sigma* triplet state.  相似文献   
92.
The newly adjusted energy-consistent nine-valence-electron pseudopotentials for K to Fr are used to calculate spectroscopic properties for the neutral and positively charged alkali dimers using coupled cluster and density functional theory. For the neutral dimers the static dipole polarizability was calculated. The coupled cluster results are all in excellent agreement with experimental values. The density functionals used can give quite different spectroscopic properties especially for the dipole polarizability, with the Perdew-Wang PW91 functional performing best.  相似文献   
93.
The ability of powdered Nylon 612 to bind methyl orange, ethyl orange, propyl orange, and butyl orange was investigated at 5, 15, 25 and 35°C in an aqueous solution. The amount of binding of the dye is much higher with this polyamide than with powdered Nylon 66 reported previously,1 although the former polymer has fewer amide end groups. The Van't Hoff plots of the first binding constant for the binding of butyl orange and propyl orange by powdered Nylon 612 exhibit a bell-shaped curve, whereas the plots for methyl orange and ethyl orange do not. Maximal binding occurs at approximately 15°C for propyl orange and at about 25°C for butyl orange. This is the first instance where the peculiar temperature dependence of the binding constant has been found in the binding of propyl orange, whose hydrophobicity is less than that of butyl orange. These tendencies can be accounted for in terms of increased hydrophobic of butyl orange. These tendencies can be accounted for in terms of increased hydrophobic domains in powdered Nylon 612 and enhanced hydrophobic contributions in the binding process.  相似文献   
94.
This paper describes a simple and convenient strategy for reducing the dimensions of organic micro-and nanostructures on metal surfaces. By varying electrochemical desorption conditions, features patterned by dip-pen nanolithography or micro contact printing and made of linear alkanethiols or selenols can be gradually desorbed in a controlled fashion. The process is referred to as electrochemical whittling because the adsorbate desorption is initiated at the exterior of the feature and moves inward as a function of time. The whittling process and adsorbate desorption were studied as a function of substrate morphology, adsorbate head and tail groups, and electrolyte solvent and salt. Importantly, one can independently address different nanostructures made of different adsorbates and effect their miniaturization based upon ajudicious selection of adsorbate, applied potential, and supporting electrolyte. Some of the physical and chemical origins of these observations have been elucidated.  相似文献   
95.
This study reports depigmenting potency of 1,3-selenazol-4-one derivatives, which would be based upon the finding of direct inhibition to mushroom tyrosinase. 1,3-Selenazol-4-one derivatives exhibited inhibitory effect on dopa oxidase activity of mushroom tyrosinase. In this study, inhibitory effects of six kinds of 1,3-selenazol-4-one derivatives (A, B, C, D, E and F) on mushroom tyrosinase were investigated. Compounds at a concentration of 500 microM exhibited 33.4-62.1% of inhibition on dopa oxidase activity of mushroom tyrosinase. Their inhibitory effects were higher than that of kojic acid (31.7%), a well known tyrosinase inhibitor. 2-(4-Methylphenyl)-1,3-selenazol-4-one (A) exhibited the strongest inhibitory effect among them dose-dependently and in competitive inhibition manner.  相似文献   
96.
Highly ordered Ni-MCM-41 samples with nearly atomically dispersed nickel ions were prepared reproducibly and characterized. Similar to the Co-MCM-41 samples, the pore diameter and porosity can be precisely controlled by changing the synthesis surfactant chain length. Nickel was incorporated by isomorphous substitution of silicon in the MCM-41 silica framework, which makes the Ni-MCM-41 a physically stable catalyst in harsh reaction conditions such as CO disproportionation to single wall carbon nanotubes or CO2 methanation. X-ray absorption spectroscopy results indicate that the overall local environment of nickel in Ni-MCM-41 was a tetrahedral or distorted tetrahedral coordination with surrounding oxygen anions. Hydrogen TPR revealed that our Ni-MCM-41 samples have high stability against reduction; however, compared to Co-MCM-41, the Ni-MCM-41 has a lower reduction temperature, and both the H2-TPR and in situ XANES TPR reveal that the reducibility of nickel is not clearly correlated with the pore radius of curvature, as in the case of Co-MCM-41. This is probably a result of nickel being thermodynamically more easily reduced than cobalt. The stability of the structural order of Ni-MCM-41 has been investigated under SWNT synthesis and CO2 methanation reaction conditions as both require catalyst exposure to reducing environments leading to formation of metallic Ni clusters. Nitrogen physisorption and XRD results show that structural order was maintained under both SWNT synthesis and CO2 methanation reaction conditions. EXAFS results demonstrate that the nickel particle size can be controlled by different prereduction temperatures but not by the pore radius of curvature as in the case of Co-MCM-41.  相似文献   
97.
Thephotochemical reaction of [3(3)](1,3,5)cyclophane 2, which is a photoprecursor for the formation of propella[3(3)]prismane 18, was studied using a sterilizing lamp (254 nm). Upon photolysis in dry and wet CH2Cl2 or MeOH in the presence of 2 mol/L aqueous HCl solution, the cyclophane 2 afforded novel cage compounds comprised of new skeletons, tetracyclo[6.3.1.0.(2,7)0(4,11)]dodeca-5,9-diene 43, hexacyclo[6.4.0.0.(2,6)0.(4,11)0.(5,10)0(9,12)]dodecane 44, and pentacyclo[6.4.0.0.(2,6)0.(4,11)0(5,10)]dodecane 45. All of these products were confirmed by the X-ray structural analyses. A possible mechanism for the formation of these photoproducts via the hexaprismane derivative 18 is proposed. The photophysical properties in the excited state of the [3n]cyclophanes ([3n]CP, n = 2-6) were investigated by measuring the emission spectra and determining the quantum yields and lifetimes of the fluorescence. All [3n]CPs show excimeric fluorescence without a monomeric one. The lifetime of the excimer fluorescence becomes gradually longer with the increasing number of the trimethylene bridges. The [3n]CPs also shows excimeric phosphorescence spectra without vibrational structures for n = 2, 4, and 5, while phosphorescence is absent for n = 3 and 6. With an increase in symmetry of the benzene skeleton in the [3(3)]- and [3(6)]CPs, the probability of the radiation (phosphorescence) process from the lowest triplet state may drastically decrease.  相似文献   
98.
Both adaptive and deleterious responses of cells to ethanol are likely triggered by short-term interactions of the cells with ethanol. Many studies have demonstrated the direct effect of ethanol on growth factor-stimulated cell proliferation. Using Swiss 3T3 cells whose growth was inhibited by ethanol in a concentration-dependent manner, we further investigated the molecular mechanisms of acute ethanol treatment by examining its effect on EGF- and PDGF-mediated cellular signaling systems for the mitogenic function. Tyrosine autophosphorylation of the growth factor receptors was partially prevented by ethanol in intact cells. When ethanol was included before or after EGF stimulation, no effect on the receptor signaling was observed. Here we also report that ethanol inhibits activation of ERK induced by both EGF and PDGF. EGF-induced JNK activation was reduced but PDGF-induced rapid JNK activation was delayed by the addition of ethanol. The balance between its inhibitory and stimulatory effect on the signaling molecules might determine the rate of cell growth.  相似文献   
99.
Mutation of the active-site residue Cys38 of N-Ada converts it from a sacrificial DNA repair protein to an enzyme that uses methanethiol as an external sacrificial reagent to repair DNA methyl phosphotriesters catalytically.  相似文献   
100.
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   
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