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.
The high‐capacity cathode material V2O5·n H2O has attracted considerable attention for metal ion batteries due to the multielectron redox reaction during electrochemical processes. It has an expanded layer structure, which can host large ions or multivalent ions. However, structural instability and poor electronic and ionic conductivities greatly handicap its application. Here, in cell tests, self‐assembly V2O5·n H2O nanoflakes shows excellent electrochemical performance with either monovalent or multivalent cation intercalation. They are directly grown on a 3D conductive stainless steel mesh substrate via a simple and green hydrothermal method. Well‐layered nanoflakes are obtained after heat treatment at 300 °C (V2O5·0.3H2O). Nanoflakes with ultrathin flower petals deliver a stable capacity of 250 mA h g?1 in a Li‐ion cell, 110 mA h g?1 in a Na‐ion cell, and 80 mA h g?1 in an Al‐ion cell in their respective potential ranges (2.0–4.0 V for Li and Na‐ion batteries and 0.1–2.5 V for Al‐ion battery) after 100 cycles.  相似文献   

2.
2-Deoxy-β-d-arabino-hexopyranose, C6H12O5, is orthorhombic, P212121, with cell dimensions at ?150° [20°], a = 6.484(2) [6.510(3)], b = 10.364(2) [10.427(4)], c = 11.134(3) [11.153(5)] Å, V = 748.2 [757.1] Å3, Z = 4, Dx = 1.457 [1.440], and Dm = [1.455] g.cm?3. The intensities of 1269 reflections were measured by using MoKα radiation. The structure was solved by direct methods, and refined by full-matrix least-squares, with anisotropic, thermal parameters for the carbon and oxygen atoms, and isotropic parameters for the hydrogen atoms. The pyranose has the 4C1(d) conformation, with puckering parameters Q = 0.563 Å, θ = 3.9°, and ? = 350.3°. The departure from ideality is very small, and less than that in β-d-glucopyranose, Q = 0.584 Å and θ = 6.9°. The β-glycosidic, CO bond is short, 1.383(4) Å, and the OCOH torsion angle is ?87°, consistent with the anomeric effect. The hydrogen-bonding scheme consists of infinite chains, with side chains terminating at a ring-oxygen atom.  相似文献   

3.
The hydrolyses of p-nitrotrifluoroacetanilide catalyzed by water and imidazole were examined at 70°C. The pH-rate constant profile of the hydrolysis in H2O was examined in the pH range 0.0–11.4. The hydrolysis was independent of pH in the region from pH 1.0 to 4.5, presumably a water-catalyzed reaction. The rate constant and the D2O solvent isotope effect for this reaction were 1.0 × 10?4 sec?1 and 3.7, respectively. Both natural imidazole and imidazolium cation catalyzed hydrolysis. The rate constant of the hydrolysis catalyzed by neutral imidazole was determined to be 5.4 × 10?3M?1 sec?1 and the D2O solvent isotope effect was 1.8.  相似文献   

4.
Proton as well as deuteron ENDOR (electron-nuclear double resonance) spectroscopy were performed of methanol dehydrogenase and pyrrolo-quinoline semiquinone (PQQH.). Samples were examined in H2O- and 2H2O-containing buffers at 4.2 °K with Ka-band (33.5 GHz) frequency. Measurements of the enzyme in 2H2O revealed that the signals observed around the proton free-precession frequency belong to exchangeable protons. Therefore, our earlier assumption (R. de Beer et al. (1979) J. Chem. Phys.70, 4491–4495) that these signals originate from protons in the aromatic ring of PQQH. is incorrect. The proton matrix signal of the enzyme in H2O and 2H2O are nearly similar, while a deuteron matrix signal is not observed in the latter case. It is concluded, therefore, that the coenzyme is situated in a hydrophobic site of the enzyme.  相似文献   

5.
Interaction between D-glucuronic acid and Zn(II), Cd(II), and Hg(II) metal ion salts has been studied in solution and solid complexes of the type M(D-glucuronate)X · nH2O and M(D-glucuronate)2·nH2O, where M = Zn(II), Cd(II), and Hg(II), X = Cl or Br, and n = 0–2 were isolated and characterized. Spectroscopic and other evidence indicated that in the metal-halide-sugar complexes the Zn(II) and Cd(II) ions bind to two D-glucuronate moieties via 06, 05 of the carboxyl oxygen atoms of the first and 04, 06' of hydroxyl and carbonyl groups of the second as well as to two H2O molecules, whereas in the corresponding M(D-glucuronate)2 · nH2O salts, the metal ions are bonded to two sugar anions through 06 and 06' of the ionized carboxyl groups and two water molecules, resulting in a six-coordination around each metal cation. The Hg(II) ion binds to 06 and 05 oxygen atoms of a sugar anion and to a halide anion or water molecule, in the Hg(D-glucuronate)X·nH2O compounds, while in the corresponding metal-glucuronate salt mercury is bonded to 06 and 06' of the two glucuronate anions with four-coordination around the Hg(II) ion. The β-anomer sugar conformation is predominant in the free acid and in these series of metal-sugar complexes.  相似文献   

6.
The reversible conformational change of DNAs and polydeoxyribonucleotides occurring before melting was followed by circular dichroism. Δθ/δT, the rate of change of ellipticity θ with temperature, was used mainly as a measure of this premelting phenomenon. If sodium ions were replaced by tetramethylammonium ions Δθ/δT decreased for poly (dA) poly (dT) and poly (dA.dT) poly (dT.dA), but increased for poly (dG.dC) poly (dC.dG). DNAs of different base composition showed no more premelting (Δθ/ΔT ~ 0) even at low molarities of TMACl provided the Na/TMA ratio was very small. For all cases studied the θ values at 0°C and at a given ionic strength were smaller in NaCl than in TMACl. When studying the series of ammonium ions from NH+4 to (C2H5)4,N+ the Δθ/ΔT values first decreased, going through zero with TMA+ io and then increased again. A tentative and qualitative explanation of our results can be given: (a) Hydration of the polymers increases in presence of TMA ions and their average stability decreases; locally, however, (AT) pairs are preferentially stabilized by TMA ions owing to a specific interaction at the level of O2 of thymine. (b) In order to explain the different behaviour of (AT) polymers and DNA, it is assumed that only the B structure is able to accommodate TMA ions in the small groove of the double stranded helix.  相似文献   

7.
The process of relaxation of energetic O ions formed via dissociative attachment of electrons to molecules in the discharge plasmas of water vapor and H2O: O2 mixtures in a strong electric field is studied by the Monte Carlo method. The probability of energetic ions being involved in threshold ion–molecular processes is calculated. It is shown that several percent of energetic O ions formed via electron attachment to H2O molecules in the course of plasma thermalization transform into OH ions via charge exchange or are destroyed with the formation of free electrons. The probabilities of charge exchange of O ions and electron detachment from them increase significantly (up to 90%) when O ions are formed via electron attachment to O2 molecules in water vapor with an oxygen additive. This effect decreases with increasing oxygen fraction in the mixture but remains appreciable even when the fraction of H2O molecules in the H2O: O2 mixture does not exceed several percent.  相似文献   

8.
In aqueous solutions of NAD(H), there is an equilibrium between two different conformations : a “folded” conformation in which adenine and nicotinamide are staked together and an “unfolded” conformation in which the two rings are without interaction.The folded conformation is the more stable in aqueous solution whereas in organic solution it is the unfolded one.As we have previously shown, the PMR spectra of Co2+—NAD(H) complexes may be related with the coenzyme conformation giving suggest to a new method for NAD(H) conformational analysis.The results of this method applied to methanol 2H2O and dioxane/2H2O solutions are reported in this paper: they are in good accordance with those of spectrofluorimetric analysis.  相似文献   

9.
《Inorganica chimica acta》2006,359(5):1619-1626
The reaction of 1,4-dimethyl-1,4,7-triazacyclononane (L-Me2) with MnCl2 · 4H2O in acetonitrile gives, in the presence of sodium formate, hydrogen peroxide, triethylamine and KPF6, the dinuclear Mn(III)–Mn(IV) complex cation [(L-Me2)2Mn2(O)2(OOCH)]2+ (1) which crystallises as the hexafluorophosphate salt.The analogous reaction with sodium benzoate, however, yields the dinuclear Mn(III)–Mn(III) complex cation [(L-Me2)2Mn2(O)(OOCC6H5)2]2+ (2), isolated also as the hexafluorophosphate salt.In the case of sodium acetate, both cations, the Mn(III)–Mn(IV) complex [(L-Me2)2Mn2(O)2(OOCCH3)]2+ (3) and the known Mn(III)–Mn(III) complex [(L-Me2)2Mn2(O)(OOCCH3)2]2+ (4) are available, depending upon the molar ratio.The single-crystal X-ray structure analyses show for the green crystals of [1][PF6]1.5[Cl]0.5 · 1.5 H2O and [3][PF6]2 · (CH3)2CO, a Mn–Mn distance of 2.620(2) and 2.628(4) Å, respectively, while for the red-violet crystal of [4][PF6]2, a Mn–Mn distance of 3.1416(8) Å is observed.All four compounds show catalytic activity for the oxidation of isopropanol with hydrogen peroxide in water and in acetonitrile to give acetone in the presence of oxalic or ascorbic acid as co-catalysts.  相似文献   

10.
《Inorganica chimica acta》1988,146(2):181-185
The reactions between [TcOCl4] and the sterically bulky thiols ArSH (Ar = 2,4,6-Me3C6H2, 2,4,6- Pri3C6H2 and 2,6-Ph2C6H3) in methanol afford complexes of formula [TcO(SAr)4] which may be isolated as salts with bulky organic cations. The molecular structure of [Bun4N][TcO(2,4,6-Me3C6H2S)4] was determined by X-ray diffraction methods. The Tc(V) centre was found to adopt the expected square pyramidal geometry in which an oxo group occupies the apical site and the four thiolate sulphurs the basal sites. The TcO distance is 1.659(11) Å and the average TcS distance 2.38(2) Å. The average cis STcS, trans STcS and OTcS angles are respectively 82.7(6)°, 138.4(3)° and 110.8(4)°.  相似文献   

11.
Abstract

Circular dichroism spectroscopy, absorption spectroscopy, measurements of Tm values, sedimentation analysis and electron microscopy were used to study properties of calf thymus DNA in methanol-water mixtures as a function of monovalent cation (Na+ or Cs+) concentration and also in the presence of divalent cations Ca2+, Mg2+, and Mn2+. In the absence of divalent cations only slight conformational changes occured and no condensation and/or aggregation could be detected. The Tm values depend on the amount of methanol and on the nature and concentration of cations. In methanol-water mixtures higher thermal stability was observed in solutions containing Cs+ ions. Up to 40% (v/v) methanol the addition of divalent ions leads to DNA stabilization. At methanol concentration higher than 50% the presence of divalent cations causes DNA condensation and denaturation even at room temperature. The denaturation is reversible with respect to EDTA addition indicating that no separation of complementary strands occured and the resulting form of DNA is probably similar to the P form. DNA destacking appears to be a direct consequence of stronger cation binding by the condensed DNA in methanol-water mixtures.  相似文献   

12.
The E0′ values for the conversion of horse heart cytochrome c from the oxidized to the reduced form as a function of temperature have been measured in 0.10 M NaCl, 0.10 M sodium phosphate, pH 7.0 solutions in H2O and D2O. In H2O, the decrease in the E0′ value is linear with increasing temperature up to 42°C. Above this temperature, the decrease is again linear but with a much greater slope. In D2O solutions, however, this biphasic behavior was not observed but instead a single line was obtained over the temperature range studied (25°C to 50°C). These results are interpreted in terms of the ability of NaCl to cause a destructuring of the bulk H2O above 42°C but not in the more stable D2O (Kreishman, Foss, Inoue and Leifer (1976) Biochemistry, 15, 5431–5435). This decrease in water structure results in a shift in the equilibrium to the larger oxidized form as indicated by the decrease in E0′.  相似文献   

13.
《Inorganica chimica acta》1988,149(1):151-154
The extraction equilibrium of the hydronium-uranium(VI)-dicyclohexano-24-crown-8 complex was carried out in the crown ether1,2-dichloroethaneHCl aqueous solution system at different temperatures. The extraction complex has the overall composition (L)2·(H3O+·χH2O)2·UO2Cl42− (L = dicyclohexano-24-crown-8). The values of the extraction equilibrium constants (Kex) increase steadily with a decrease in temperature: 13.5 (298 K), 7.96 (301 K), 4.20 (303 K) and 2.07 (305 K). A plot of log Kex against 1/T shows a straight line. The value of the enthalpy change, ΔH°, was calculated from the slope and equals −212 kJ mol−1. The value of the entropy change, ΔS°, was calculated from ΔH° and Kex and equals −690 J K−1 mol−1, whereas ΔG° = −6.45 kJ mol−1. Comparing these thermodynamic parameters with those of the dicyclohexano-18-crown-6 isomer A [1] (ΔS° = −314 J K−1 mol−1, ΔH° = −101 kJ mol−1 and ΔG° = −8.37 kJ mol−1), it can be seen that ΔH° and ΔS° are more negative for the former than for the latter, and both are enthalpy-stabilized complexes. The molecular structure of the complex has the feature that there are two H5O2+ ions in it, in contrast to the H3O+ ions in the dicyclohexano-18-crown-6 isomer A complex [1]. Each of the H5O2+ ions is held in the crown ether cavity by four hydrogen bonds. The H5O2+ ion has a central bond. The uranium atom forms UO2Cl42− as a counterion away from the crown ether. The formation of this complex is in good agreement with more negative entropy change and less negative free energy change, as mentioned above.  相似文献   

14.
Abstract

The effects of Ca2+ ions on 3H-RO 5–4864 binding to the peripheral benzodiazepine receptor were examined. Preincubation of rat kidney membranes with Ca2+ at 37°C produced a dose-dependent inhibition of 3H-RO 5–4864 binding. No inhibition was observed in membranes preincubated at 0°C.

The effect of Ca2+ was competitive in nature and was fully reversed by the addition of EGTA. At 1 mM, the maximal effect was achieved with CaCl2, whereas CoCl2 and CdCl2 had lesser effects. No other divalent cation salts examined decreased 3H-RO 5–4864 binding to rat kidney membranes. Collectively, these data demonstrate that the affinity of 3H-RO 5–4864 binding to rat kidney membranes is regulated by Ca2+ and suggest the presence of cation recognition binding sites coupled to the peripheral benzodiazepine receptor.  相似文献   

15.
Proton and phosphorus-31 nuclear spin–lattice relaxation times T1 and spin–spin relaxation times T2 have been measured on the single-stranded polyriboadenylic acid [poly(A)]–Mn2+ system in a neutral D2O solution in the temperature range 10°–90°C at 100 and 40.5 MHz, respectively, with the Fourier transform nmr method. Minimum values of T1 have been found for all these nuclei, which have enabled the exact estimation of apparent distances from Mn2+ to H2, H8, H1′, and the phosphorus nucleus to be 4.7, 4.1, 5.2, and 3.0 Å, respectively. The electron spin of Mn2+ penetrates into the phosphorus nucleus, giving 31P hyperfine coupling of more than 106 Hz. Evidence of penetration of the electron spin into H8 and H2 is also obtained, suggesting direct coordination of nitrogen atoms of the adenine ring to the Mn2+ Ion. Combined with the result from proton relaxation enhancement of water, it is concluded that every Mn2+ ion added is bound directly to two phosphate groups with a Mn2+–phosphorus distance of 3.3 Å, while a part of the Mn2+ ions are simultaneouly bound to the adenine ring. It is estimated that 39 ± 13% and 13 ± 5% of Mn2+ are coordinated by N7 and N3 (or N1), respectively. The motional freedom of poly(A) in the environment of the Mn2+ binding site has been found to be quenched to the extent that the rotational motion becomes several times slower than that of the corresponding Mn2+–free poly(A). The activation energies for the molecular motion are, however, practically unchanged from those for Mn2+–free poly(A), and are found to be 8.3, 8.5, 6.1, and 8.7 kcal/mol for H8, H2, H1′, and phosphorus, respectively. T2 of phosphorus is determined by the dissociation rate (k?1) of Mn2+ from the phosphate group for the whole temperature range studied with activation enthalpy of 6.5 kcal/mol. The dissociation rates of Mn2+ from the adenine ring are also estimated from proton T2 values below 50°C.  相似文献   

16.
The CH3 + ion, formed in ionized methane, undergoes consecutive eliminative condensation reactions with methane to form the carbonium ions C2H5 +, i-C3H7 + and t-C4H9 +. AtT<500°K, \(N_{CH_4 } \) ?1016 cm?3 these ions react with NH3 in competitive condensation-H+ transfer reactions, e.g. $$\begin{gathered} C_2 H_5 ^ + + NH_3 \xrightarrow{M} C_2 H_5 NH_3 ^ + \hfill \\ - - - \to NH_4 ^ + + C_2 H_4 \hfill \\ \end{gathered} $$ At particle densities of \(N_{CH_4 } \) <1016 cm?3 proton transfer is the only significant reaction channel. At \(N_{CH_4 } \) >1017 cm?3 condensation constitutes 5–20% of the overall reactions. The product of the condensation reaction further associates with CO2 to form C2H5NH3 +·CO2; the atomic composition of this cluster ion is identical with the protonated amino acid alanine. The carbonium ions i-C3H7 + and t-C4H9 + condense also with HCN to yield protonated isocyanides. HCNH+ also appears to condense with HCN atT>570°K, and form cluster ions with HCN at lower temperatures. The rate constants of the condensation reactions vary with temperature and pressure in a complex manner. Under conditions similar to those on Titan at an altitude of 100 km (T=100–150°K, \(N_{CH_4 } \) ≈1018 cm?3), with a methane atmosphere containing 1% H2 and traces of NH3 and H2O, ion-molecule condensation reactions followed by H+ transfer are expected to lead to the atmospheric synthesis of C2H6, C3H8, CH3OH, C2H5OH and the terminal ions NH4 +, CH3NH3 + and C2H5NH3 +. At higher temperatures (250°K<T<400°K), the synthesis of i-C4H10, i-C3H7OH and t-C4H9OH and of the ions i-C3H7NH3 + and t-C4H9NH3 + is also expected. Electron recombination of the terminal ions may yield amines, imines and nitriles. Cycles of protonation and dissociative recombination of the alkanes and alcohols produced in condensation reactions will also produce unsaturated hydrocarbons, ketones and aldehydes in the ionized atmosphere.  相似文献   

17.
A new tetracopper(II) complex bridged both by oxamido and carboxylato groups, namely [Cu4(dmaepox)2(bpy)2](NO3)2·2H2O, where H3dmaepox and bpy represent N‐benzoato‐N′‐ (3‐methylaminopropyl)oxamide and 2,2′‐bipyridine, was synthesized, and its structure reveals the presence of a centrosymmetric cyclic tetracopper(II) cation assembled by a pair of cis‐dmaepox3–‐ bridged dicopper(II) units through the carboxylato groups, in which the endo‐ and exo‐copper(II) ions bridged by the oxamido group have a square‐planar and a square‐pyramidal coordination geometries, respectively. The aromatic packing interactions assemble the complex molecules to a two‐dimensional supramolecular structure. The reactivity toward DNA and protein bovine serum albumin (BSA) indicates that the complex can interact with herring sperm DNA through the intercalation mode and the binding affinity is dominated by the hydrophobicity and chelate ring arrangement around copper(II) ions and quenches the intrinsic fluorescence of BSA via a static process. The cytotoxicity of the complex shows selective cancer cell antiproliferative activity.  相似文献   

18.
Density functional calculations have been used to investigate the interactions of 1-(2-hydroxyethyl)-3-methylimidazolium ([C2OHmim]+)-based ionic liquids (hydroxyl ILs) with water (H2O), methanol (CH3OH), and dimethyl sulfoxide (DMSO). It was found that the cosolvent molecules interact with the anion and cation of each ionic liquid through different atoms, i.e., H and O atoms, respectively. The interactions between the cosolvent molecules and 1-ethyl-3-methylimizolium ([C2mim]+)-based ionic liquids (nonhydroxyl ILs) were also studied for comparison. In the cosolvent–[nonhydroxyl ILs] systems, a furcated H-bond was formed between the O atom of the cosolvent molecule and the C2-H and C6-H, while there were always H-bonds involving the OH group of the cation in the cosolvent–[hydroxyl ILs] systems. Introducing an OH group on the ethyl side of the imidazolium ring may change the order of solubility of the molecular liquids.  相似文献   

19.
The 18O content of CO2 is a powerful tracer of photosynthetic activity at the ecosystem and global scale. Due to oxygen exchange between CO2 and 18O-enriched leaf water and retrodiffusion of most of this CO2 back to the atmosphere, leaves effectively discriminate against 18O during photosynthesis. Discrimination against 18O ( Δ 18O) is expected to be lower in C4 plants because of low ci and hence low retrodiffusing CO2 flux. C4 plants also generally show lower levels of carbonic anhydrase (CA) activities than C3 plants. Low CA may limit the extent of 18O exchange and further reduce Δ 18O. We investigated CO2–H2O isotopic equilibrium in plants with naturally low CA activity, including two C4 (Zea mays, Sorghum bicolor) and one C3 (Phragmites australis) species. The results confirmed experimentally the occurrence of low Δ 18O in C4, as well as in some C3, plants. Variations in CA activity and in the extent of CO2–H2O isotopic equilibrium ( θ eq) estimated from on-line measurements of Δ 18O showed large range of 0–100% isotopic equilibrium ( θ eq = 0–1). This was consistent with direct estimates based on assays of CA activity and measurements of CO2 concentrations and residence times in the leaves. The results demonstrate the potential usefulness of Δ 18O as indicator of CA activity in vivo. Sensitivity tests indicated also that the impact of θ eq < 1 (incomplete isotopic equilibrium) on 18O of atmospheric CO2 can be similar for C3 and C4 plants and in both cases it increases with natural enrichment of 18O in leaf water.  相似文献   

20.
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