首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
A mononuclear‐cobalt(II)‐substituted silicotungstate, K10[Co(H2O)2(γ‐SiW10O35)2] ? 23 H2O (POM‐ 1 ), has been evaluated as a light‐driven water‐oxidation catalyst. With in situ photogenerated [Ru(bpy)3]3+ (bpy=2,2′‐bipyridine) as the oxidant, quite high catalytic turnover number (TON; 313), turnover frequency (TOF; 3.2 s?1), and quantum yield (ΦQY; 27 %) for oxygen evolution at pH 9.0 were acquired. Comparison experiments with its structural analogues, namely [Ni(H2O)2(γ‐SiW10O35)2]10? (POM‐ 2 ) and [Mn(H2O)2(γ‐SiW10O35)2]10? (POM‐ 3 ), gave the conclusion that the cobalt center in POM‐ 1 is the active site. The hydrolytic stability of the title polyoxometalate (POM) was confirmed by extensive experiments, including UV/Vis spectroscopy, linear sweep voltammetry (LSV), and cathodic adsorption stripping analysis (CASA). As the [Ru(bpy)3]2+/visible light/sodium persulfate system was introduced, a POM–photosensitizer complex formed within minutes before visible‐light irradiation. It was demonstrated that this complex functioned as the active species, which remained intact after the oxygen‐evolution reaction. Multiple experimental parameters were investigated and the catalytic activity was also compared with the well‐studied POM‐based water‐oxidation catalysts (i.e., [Co4(H2O)2(α‐PW9O34)2]10? (Co4‐POM) and [CoIIICoII(H2O)W11O39]7? (Co2‐POM)) under optimum conditions.  相似文献   

2.
The development of visible‐light‐induced photocatalysts for chemoselective functional group transformations has received considerable attention. Polyoxometalates (POMs) are potential materials for efficient photocatalysts because their properties can be precisely tuned by changing their constituent elements and structures and by the introduction of additional metal cations. Furthermore, they are thermally and oxidatively more stable than the frequently utilized organometallic complexes. The visible‐light‐responsive tetranuclear cerium(III)‐containing silicotungstate TBA6[{Ce(H2O)}2{Ce(CH3CN)}24‐O)(γ‐SiW10O36)2] (CePOM; TBA=tetra‐n‐butylammonium) has now been synthesized; when CePOM was irradiated with visible light (λ>400 nm), a unique intramolecular CeIII‐to‐POM(WVI) charge transfer was observed. With CePOM, the photocatalytic oxidative dehydrogenation of primary and secondary amines as well as the α‐cyanation of tertiary amines smoothly proceeded in the presence of O2 (1 atm) as the sole oxidant.  相似文献   

3.
The crystal structure of methyl α‐d ‐mannopyranosyl‐(1→3)‐2‐O‐acetyl‐β‐d ‐mannopyranoside monohydrate, C15H26O12·H2O, ( II ), has been determined and the structural parameters for its constituent α‐d ‐mannopyranosyl residue compared with those for methyl α‐d ‐mannopyranoside. Mono‐O‐acetylation appears to promote the crystallization of ( II ), inferred from the difficulty in crystallizing methyl α‐d ‐mannopyranosyl‐(1→3)‐β‐d ‐mannopyranoside despite repeated attempts. The conformational properties of the O‐acetyl side chain in ( II ) are similar to those observed in recent studies of peracetylated mannose‐containing oligosaccharides, having a preferred geometry in which the C2—H2 bond eclipses the C=O bond of the acetyl group. The C2—O2 bond in ( II ) elongates by ~0.02 Å upon O‐acetylation. The phi (?) and psi (ψ) torsion angles that dictate the conformation of the internal O‐glycosidic linkage in ( II ) are similar to those determined recently in aqueous solution by NMR spectroscopy for unacetylated ( II ) using the statistical program MA′AT, with a greater disparity found for ψ (Δ = ~16°) than for ? (Δ = ~6°).  相似文献   

4.
A combination of electrospray ionisation (ESI), multistage and high‐resolution mass spectrometry experiments is used to examine the gas‐phase fragmentation reactions of the three isomeric phenylalanine derivatives, α‐phenylalanine, β2‐phenylalanine and β3‐phenylalanine. Under collision‐induced dissociation (CID) conditions, each of the protonated phenylalanine isomers fragmented differently, allowing for differentiation. For example, protonated β3‐phenylalanine fragments almost exclusively via the loss of NH3, only β2‐phenylalanine via the loss of H2O, while α‐ and β2‐phenylalanine fragment mainly via the combined losses of H2O + CO. Density functional theory (DFT) calculations were performed to examine the competition between NH3 loss and the combined losses of H2O and CO for each of the protonated phenylalanine isomers. Three potential NH3 loss pathways were studied: (i) an aryl‐assisted neighbouring group; (ii) 1,2 hydride migration; and (iii) neighbouring group participation by the carboxyl group. Finally, we have shown that isomer differentiation is also possible when CID is performed on the protonated methyl ester and methyl amide derivatives of α‐, β2‐ and β3‐phenylalanines. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

5.
The title compound, C32H45N2O+·Br?·0.5H2O, has the outer two six‐membered rings in chair conformations, while the central ring is in an 8β,9α‐half‐chair conformation. The five‐mem­bered ring of the steroid nucleus adopts a slightly deformed 14α‐envelope conformation. The pyridyl­methyl­ene moiety has an E configuration with respect to the hydroxyl group at position 17. The structure is stabilized by a network of O—H?Br‐type intermolecular hydrogen bonds.  相似文献   

6.
The crystal structure of octaguanidinium α‐silicodiplatino­decatungstate hexahydrate, (CH6N3)8[α‐SiPt2W10O40]·6H2O, has been analyzed via a high‐energy X‐ray diffraction experiment at the SPring‐8 BL04B2 beamline. The title compound contains a novel α‐Keggin heteropolyanion in which two of the addenda atoms are replaced by Pt atoms. W and Pt atoms occupy the same coordinates; the occupancy fractions are (W) and (Pt), and the α‐Keggin anion has symmetry. The two types of W(Pt)—W(Pt) distance are in the ranges 3.3565 (4)–3.3704 (4) and 3.7033 (4)–3.7100 (4) Å, the four types of W(Pt)—O bond length are in the ranges 1.721 (5)–1.725 (5), 1.910 (5)–1.932 (5), 1.934 (5)–1.956 (5) and 2.339 (4)–2.348 (4) Å, and the Si—O bond length is 1.646 (4) Å.  相似文献   

7.
The new supramolecular compound [H2bpp][{Cu(Hbpy)2}{α‐HP2W18O62}]·4H2O ( 1 ) (bpy = 4,4′‐bipyridine, bpp = 1,3‐bis(4‐pyridyl)propane) was synthesized hydrothermally and characterized byelemental analysis, IR spectroscopy, thermogravimetric analysis and single‐crystal X‐ray diffraction. In compound 1 , the cationic fragment [Cu(Hbpy)2]+ connects to the Dawson anion through a coordinating Cu←O bond, and the copper atom is coordinated by another polyoxoanion through a weak covalent bond with a Cu1–O26 distance of 2.879(2) Å, forming a polymeric chain. The bpy ligand in [Cu(Hbpy)2]+ adopts a monodentate coordination mode, the other nitrogen atom of the bpy ligand is protonated. The protonated Hbpy+ acts as hydrogen‐bond donor and constructs a two‐dimensional double‐sheet supramolecular network involving the one‐dimensional chains through the hydrogen bonds. The H2bpp2+ ion connects twoα‐HP2W18O626– clusters from two supramolecular networks through hydrogen bonds and creates a three‐dimensional supramolecular architecture. The thermal decomposition of 1 happens over a wide temperature range (450–800 °C), which indicates that it might include complicated oxidation–reduction processes.  相似文献   

8.
Al‐ and Ga‐containing open‐Dawson polyoxometalates (POMs), K10[{Al4(μ‐OH)6}{α,α‐Si2W18O66}] · 28.5H2O ( Al4 ‐ open ) and K10[{Ga4(μ‐OH)6}(α,α‐Si2W18O66)] · 25H2O ( Ga4 ‐ open ) were synthesized by the reaction of trilacunary Keggin POM, [A‐α‐SiW9O34]10–, with Al(NO3)3 · 9H2O or Ga(NO3)3 · nH2O, and unequivocally characterized by single‐crystal X‐ray analysis, 29Si and 183W NMR, and FT‐IR spectroscopy as well as elemental analysis and TG/DTA. Single‐crystal X‐ray analysis revealed that the {M4(μ‐OH)6}6+ (M = Al, Ga) clusters were included in an open pocket of the open‐Dawson polyanion, [α,α‐Si2W18O66]16–, which was constituted by the fusion of two trilacunary Keggin POMs via two W–O–W bonds. These two open‐Dawson structural POMs showed clear difference of the bite angles depending on the size of ionic radii. In cases of both compounds, the solution 29Si and 183W NMR spectra in D2O showed only one signal and five signals, respectively. These spectra were consistent with the molecular structures of Al4 ‐ and Ga4 ‐ open , suggesting that these polyoxoanions were obtained as single species and maintained their molecular structures in solution.  相似文献   

9.
α‐Fe2O3 nanoparticles are uniformly coated on the surface of α‐MoO3 nanorods through a two‐step hydrothermal synthesis method. As the anode of a lithium‐ion battery, α‐Fe2O3@α‐MoO3 core–shell nanorods exhibit extremely high lithium‐storage performance. At a rate of 0.1 C (10 h per half cycle), the reversible capacity of α‐Fe2O3@α‐MoO3 core–shell nanorods is 1481 mA h g?1 and a value of 1281 mA h g?1 is retained after 50 cycles, which is much higher than that retained by bare α‐MoO3 and α‐Fe2O3 and higher than traditional theoretical results. Such a good performance can be attributed to the synergistic effect between α‐Fe2O3 and α‐MoO3, the small size effect, one‐dimensional nanostructures, short paths for lithium diffusion, and interface spaces. Our results reveal that core–shell nanocomposites have potential applications as high‐performance lithium‐ion batteries.  相似文献   

10.
In the title compounds, C18H20N2O2, (I), and C14H11N3O4·0.5H2O, (II), respectively, the oxime groups have an E configuration. In (I), the mol­ecules exist as polymers bound by intermolecular C—H⋯O and O—H⋯N hydrogen bonds around inversion centres. In (II), intermolecular OW—H⋯N, OW—H⋯O and O—H⋯OW interactions stabilize the molecular packing.  相似文献   

11.
The X‐ray analyses of 2,3,4,6‐tetra‐O‐acetyl‐α‐d ‐glucopyranosyl fluoride, C14H19FO9, (I), and the corresponding maltose derivative 2,3,4,6‐tetra‐O‐acetyl‐α‐d ‐glucopyranosyl‐(1→4)‐2,3,6‐tri‐O‐acetyl‐α‐d ‐glucopyranosyl fluoride, C26H35FO17, (II), are reported. These add to the series of published α‐glycosyl halide structures; those of the peracetylated α‐glucosyl chloride [James & Hall (1969). Acta Cryst. A 25 , S196] and bromide [Takai, Watanabe, Hayashi & Watanabe (1976). Bull. Fac. Eng. Hokkaido Univ. 79 , 101–109] have been reported already. In our structures, which have been determined at 140 K, the glycopyranosyl ring appears in a regular 4C1 chair conformation with all the substituents, except for the anomeric fluoride (which adopts an axial orientation), in equatorial positions. The observed bond lengths are consistent with a strong anomeric effect, viz. the C1—O5 (carbohydrate numbering) bond lengths are 1.381 (2) and 1.381 (3) Å in (I) and (II), respectively, both significantly shorter than the C5—O5 bond lengths, viz. 1.448 (2) Å in (I) and 1.444 (3) Å in (II).  相似文献   

12.
In light of the serious challenge of severe global energy shortages, p‐type dye‐sensitized solar cells (p‐DSSCs) have attracted increasing levels of interest. The potential of three Keggin‐type transition metal‐substituted polyoxometalates, TBA8Na2[SiW9O37{Co(H2O)3}]? 11 H2O (SiW9Co3), TBA4[(SiO4)W10MnIII2O36H6]?1.5 CH3CN? 2 H2O (SiW10MnIII2), and TBA3.5H5.5[(SiO4)W10MnIII/IV2O36]? 10 H2O?0.5 CH3CN (SiW10MnIII/IV2) has been explored as pure inorganic dye photosensitizers for p‐DSSCs (TBA=(n‐C4H9)4N+). The three dyes show overall conversion efficiencies of 0.038, 0.029, and 0.027 %, respectively, all of which are higher than that of coumarin 343 (0.017 %). These polyoxometalates are the first three pure inorganic dyes reported for use with p‐DSSCs and therefore demonstrate a new strategy for designing efficient dyes, especially pure inorganic dyes. Moreover, they broaden the range of applications for polyoxometalates.  相似文献   

13.
A new inorganic–organic hybrid based on an aspartate functionalized polyoxomolybdate, [pentaaquacobalt(II)]‐μ‐aspartate‐[γ‐octamolybdate]‐μ‐aspartate‐[pentaaquacobalt(II)] tetrahydrate, [Co2(C4H6NO4)2(γ‐Mo8O26)(H2O)10]·4H2O ( 1 ), has been synthesized under hydrothermal conditions from the reaction of an Evans–Showell‐type polyoxometalate, (NH4)6[Co2Mo10H4O38], and l ‐aspartic acid. The complex exhibits a supramolecular three‐dimensional framework structure in the crystal lattice. Compound 1 was structurally characterized by elemental analyses, IR and UV–Vis (diffuse reflectance) spectroscopy and single‐crystal X‐ray diffraction. In this compound, aspartic acid acts as a bridge between the two Co atoms and the Mo centres, with the –CH2COOH side chain directly linked to the Mo centre in γ‐[Mo8O26]4? and the α‐carboxylate side chain bound to the Co centre. Commonly, the binding of transition‐metal complexes to POMs involves coordination of the metal to a terminal O atom of the POM so that 1 , with a bridging ligand between Mo and Co atoms, belongs to a separate class of hybrid materials. While the starting materials are both chiral and one might expect them to form a chiral hybrid, the decomposition of the chiral Evans–Showell‐type POM and its conversion to the centrosymmetric γ‐octamolybdate POM, plus the presence of two aspartate ligands centrosymmetrically placed on either side of the POM, leads to the formation of an achiral hybrid. We have studied energetically by means of density functional theory (DFT) calculations and using the Bader's `atoms‐in‐molecules' analysis the electrostatically enhanced hydrogen bonds (EEHBs) observed in the solid state of 1 , which are crucial for the formation of one‐dimensional supramolecular assemblies.  相似文献   

14.
The α‐[P2W18O62]6–‐based coordination polymer [Cu2(phen)3(H2O)3(P2W18O62)][Cu(phen)2(H2O)] · 5H2O ( 1 ) (phen = phenanthroline), was hydrothermally synthesized and characterized by single‐crystal and powder X‐ray diffraction, IR and UV/Vis diffuse reflection spectroscopy, and elemental analysis. Structural studies reveal that compound 1 exhibits a three‐dimensional (3D) supramolecular structure based on π–π and hydrogen bonding interactions. In addition, visible light driven photocatalytic experiments of compound 1 were also studied.  相似文献   

15.
New heteroatom polyoxovanadates (POVs) were synthesized by applying a water‐soluble high‐nuclearity cluster as new synthon. The [V15Sb6O42]6? cluster shell exhibiting D3 symmetry was in situ transformed into completely different cluster shells, namely, the α‐[V14Sb8O42]4? isomer with D2d and the β‐[V14Sb8O42]4? isomer with D2h symmetry. The solvothermal reaction of {Ni(en)3}3[V15Sb6O42(H2O)x] ? 15 H2O (x=0 or 1; en=ethylenediamine) in water led to the crystallization of [{Ni(en)2}2V14Sb8O42] ? 5.5 H2O containing the β‐isomer. The addition of [Ni(phen)3](ClO4)2 ? 0.5 H2O (phen=1,10‐phenanthroline) to the reaction slurry gave the new compound {Ni(phen)3}2[V14Sb8O42] ? phen ? 12 H2O with the α‐isomer. Both transformation reactions are complex due the change of symmetry, the chemical composition, and rearrangement of the VO5 square pyramids and Sb2O5 handle‐like moieties.  相似文献   

16.
Two Vanadium‐substituted Keggin‐type polyoxometalates, K3H2[α‐SiVW11O40]?6H2O (SiVW11) and K4H2[γ(1, 2)‐SiV2W10O40]?4H2O (SiV2W10) were first successfully immobilized on 4‐aminobenzoic acid modified glass carbon electrodes respectively by layer‐by‐layer assembly with poly (ethylenimine) (PEI) as counterions. The regular growth processes were monitored by cyclic voltammetry (CV), and it was proved that the multilayer films were uniform and stable. The cyclic voltammetry results indicated that the electrochemical behavior of two multilayer films was similar, and their redox couples are pH‐ and scan rate‐dependent. The multilayer films show favorable electrocatalytic active toward the reduction of NO2?, IO3? and H2O2.  相似文献   

17.
Four structures of oxoindolyl α‐hydroxy‐β‐amino acid derivatives, namely, methyl 2‐{3‐[(tert‐butoxycarbonyl)amino]‐1‐methyl‐2‐oxoindolin‐3‐yl}‐2‐methoxy‐2‐phenylacetate, C24H28N2O6, (I), methyl 2‐{3‐[(tert‐butoxycarbonyl)amino]‐1‐methyl‐2‐oxoindolin‐3‐yl}‐2‐ethoxy‐2‐phenylacetate, C25H30N2O6, (II), methyl 2‐{3‐[(tert‐butoxycarbonyl)amino]‐1‐methyl‐2‐oxoindolin‐3‐yl}‐2‐[(4‐methoxybenzyl)oxy]‐2‐phenylacetate, C31H34N2O7, (III), and methyl 2‐[(anthracen‐9‐yl)methoxy]‐2‐{3‐[(tert‐butoxycarbonyl)amino]‐1‐methyl‐2‐oxoindolin‐3‐yl}‐2‐phenylacetate, C38H36N2O6, (IV), have been determined. The diastereoselectivity of the chemical reaction involving α‐diazoesters and isatin imines in the presence of benzyl alcohol is confirmed through the relative configuration of the two stereogenic centres. In esters (I) and (III), the amide group adopts an anti conformation, whereas the conformation is syn in esters (II) and (IV). Nevertheless, the amide group forms intramolecular N—H...O hydrogen bonds with the ester and ether O atoms in all four structures. The ether‐linked substituents are in the extended conformation in all four structures. Ester (II) is dominated by intermolecular N—H...O hydrogen‐bond interactions. In contrast, the remaining three structures are sustained by C—H...O hydrogen‐bond interactions.  相似文献   

18.
The reaction of Na10[A-α-SiW9O34] · 18H2O with CuCl2 · 2H2O in the participation of ethylenediamine (En) under hydrothermal conditions resulted in a 2D organic-inorganic hybrid monocopper(II)-substituted Keggin silicotungstate [Cu(En)2(H2O)]2[Cu(En)2]4[Si2Cu2W22O78] · 7H2O (I), which was structurally characterized by elemental analyses, IR spectrum, UV spectrum, powder X-ray diffraction (PXRD), and single-crystal X-ray diffraction. Single crystal structural analysis shows that adjacent monocopper(II)-substituted Keggin silicotungstate [Si2Cu2W22O78]12? dimeric subunits are interconnected by sharing terminal oxygen atoms to make the 1D polymeric linear chain and neighboring chains are combined with each other through [Cu(En)2]2+ connectors giving rise to an interesting 2D organic-inorganic hybrid sheet architecture with a 4-connected topology. To our knowledge, I is the rare organic-inorganic hybrid 2D polyoxometate constructed by mono-transition-metal substituted Keggin silicotungstate dimeric subunits. The photocatalytic measurement illustrates that I can to some extent inhibit the photodegradation of rhodamine-B.  相似文献   

19.
α‐Oxo­benzene­acetic (phenyl­glyoxy­lic) acid, C8H6O3, adopts a transoid di­carbonyl conformation in the solid state, with the carboxyl group rotated 44.4 (1)° from the nearly planar benzoyl moiety. The heterochiral acid‐to‐ketone catemers [O?O = 2.686 (3) and H?O = 1.78 (4) Å] have a second, longer, intermolecular O—H?O contact to a carboxyl sp3 O atom [O?O = 3.274 (2) and H?O = 2.72 (4) Å], with each flat ribbon‐like chain lying in the bc plane and extending in the c direction. In α‐oxo‐2,4,6‐tri­methyl­benzene­acetic (mesityl­glyoxy­lic) acid, C11H12O3, the ketone is rotated 49.1 (7)° from planarity with the aryl ring and the carboxyl group is rotated a further 31.2 (7)° from the ketone plane. The solid consists of chiral conformers of a single handedness, aggregating in hydrogen‐bonding chains whose units are related by a 31 screw axis, producing hydrogen‐bonding helices that extend in the c direction. The hydrogen bonding is of the acid‐to‐acid type [O?O = 2.709 (6) and H?O = 1.87 (5) Å] and does not formally involve the ketone; however, the ketone O atom in the acceptor mol­ecule has a close polar contact with the same donor carboxyl group [O?O = 3.005 (6) and H?O = 2.50 (5) Å]. This secondary hydrogen bond is probably a major factor in stabilizing the observed cisoid di­carbonyl conformation. Several intermolecular C—H?O close contacts were found for the latter compound.  相似文献   

20.
Depsipeptides and cyclodepsipeptides are analogues of the corresponding peptides in which one or more amide groups are replaced by ester functions. Reports of crystal structures of linear depsipeptides are rare. The crystal structures and conformational analyses of four depsipeptides with an alternating sequence of an α,α‐disubstituted α‐amino acid and an α‐hydroxy acid are reported. The molecules in the linear hexadepsipeptide amide in (S)‐Pms‐Acp‐(S)‐Pms‐Acp‐(S)‐Pms‐Acp‐NMe2 acetonitrile solvate, C47H58N4O9·C2H3N, ( 3b ), as well as in the related linear tetradepsipeptide amide (S)‐Pms‐Aib‐(S)‐Pms‐Aib‐NMe2, C28H37N3O6, ( 5a ), the diastereoisomeric mixture (S,R)‐Pms‐Acp‐(R,S)‐Pms‐Acp‐NMe2/(R,S)‐Pms‐Acp‐(R,S)‐Pms‐Acp‐NMe2 (1:1), C32H41N3O6, ( 5b ), and (R,S)‐Mns‐Acp‐(S,R)‐Mns‐Acp‐NMe2, C30H37N3O6, ( 5c ) (Pms is phenyllactic acid, Acp is 1‐aminocyclopentanecarboxylic acid and Mns is mandelic acid), generally adopt a β‐turn conformation in the solid state, which is stabilized by intramolecular N—H…O hydrogen bonds. Whereas β‐turns of type I (or I′) are formed in the cases of ( 3b ), ( 5a ) and ( 5b ), which contain phenyllactic acid, the torsion angles for ( 5c ), which incorporates mandelic acid, indicate a β‐turn in between type I and type III. Intermolecular N—H…O and O—H…O hydrogen bonds link the molecules of ( 3a ) and ( 5b ) into extended chains, and those of ( 5a ) and ( 5c ) into two‐dimensional networks.  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