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.
We have previously shown that methionine–heme iron coordination is perturbed in domain-swapped dimeric horse cytochrome c. To gain insight into the effect of methionine dissociation in dimeric cytochrome c, we investigated its interaction with cyanide ion. We found that the Soret and Q bands of oxidized dimeric cytochrome c at 406.5 and 529 nm redshift to 413 and 536 nm, respectively, on addition of 1 mM cyanide ion. The binding constant of dimeric cytochrome c and cyanide ion was obtained as 2.5 × 104 M?1. The Fe–CN and C–N stretching (ν Fe–CN and ν CN) resonance Raman bands of CN?-bound dimeric cytochrome c were observed at 443 and 2,126 cm?1, respectively. The ν Fe–CN frequency of dimeric cytochrome c was relatively low compared with that of other CN?-bound heme proteins, and a relatively strong coupling between the Fe–C–N bending and porphyrin vibrations was observed in the 350–450-cm?1 region. The low ν Fe–CN frequency suggests weaker binding of the cyanide ion to dimeric cytochrome c compared with other heme proteins possessing a distal heme cavity. Although the secondary structure of dimeric cytochrome c did not change on addition of cyanide ion according to circular dichroism measurements, the dimer dissociation rate at 45 °C increased from (8.9 ± 0.7) × 10?6 to (3.8 ± 0.2) × 10?5 s?1, with a decrease of about 2 °C in its dissociation temperature obtained with differential scanning calorimetry. The results show that diatomic ligands may bind to the heme iron of dimeric cytochrome c and affect its stability.  相似文献   

2.
Polybenzimidazoles (PBIs) are promising materials to replace Nafion as the electrolyte in polymer electrolyte membrane fuel cells (PEMFCs). The challenge with these materials is to achieve a good compromise between the H3PO4 doping level and membrane stability. This can be obtained by a proper monomer design, which can lead to better performing membrane electrode assemblies (MEAs), in terms of durability, acid leaching, and electrode safety. Here the easy and inexpensive synthesis of hexafluoropropylidene oxyPBI (F6‐oxyPBI) and bisulfonated hexafluoropropylidene oxyPBI (F6‐oxyPBI‐2SO3H) is reported. The membranes based on F6‐oxyPBI‐2SO3H are more stable in an oxidative environment and more mechanically resistant than standard PBI and F6‐oxyPBI. Whereas the attainable doping levels are low because of fluorine‐induced hydrophobicity, polysulfonation allows high proton conductivity, and fuel cell performances better than those reported for MEAs with F6PBI‐ or PBI membranes with much higher doping levels. In the case of MEA with a F6‐oxyPBI‐2SO3H membrane, a peak power density of 360 mW cm?2 is measured. Fuel cell performances of 604 mV at 0.2 A cm?2 are maintained for 800 h without membrane degradation. Low H2 permeability is measured, which remains almost unaffected during a 1000 h life‐test.  相似文献   

3.
The B3LYP/6-31G (d) method of density functional theory (DFT) was used to study molecular geometry, electronic structure, infrared spectrum (IR) and thermodynamic properties. The heat of formation (HOF) and calculated density were estimated to evaluate the detonation properties using Kamlet–Jacobs equations. Thermal stability of 3,5,7,10,12,14,15,16-octanitro- 3,5,7,10,12,14,15,16-octaaza-heptacyclo[7.5.1.12,8.01,11.02,6.04,13.06,11]hexadecane (cage-tetranitrotetraazabicyclooctane) was investigated by calculating the bond dissociation energy (BDE) at unrestricted B3LYP/6-31G (d) level. The calculated results show that the N–NO2 bond is a trigger bond during thermolysis initiation process. The crystal structure obtained by molecular mechanics (MM) methods belongs to Pna21 space group, with cell parameters a?=?12.840 Å, b?=?9.129 Å, c?=?14.346 Å, Z?=?6 and ρ?=?2.292 g·cm?3. Both the detonation velocity of 9.96 km·s?1 and the detonation pressure of 47.47 GPa are better than those of CL-20. According to the quantitative standard of energetics and stability, as a high energy density compound (HEDC), cage-tetranitrotetraazabicyclooctane essentially satisfies this requirement.  相似文献   

4.
NAD+-dependent formate dehydrogenase (FDH, EC 1.2.1.2) is of use in the regeneration of NAD(P)H coenzymes, and therefore has strong potential for practical application in chemical and medical industries. A low-cost production of recombinant Escherichia coli (E. coli) containing FDH from Candida methylica (cmFDH) was optimized in molasses-based medium by using response surface methodology (RSM) based on central composite design (CCD). The beet molasses as a sole carbon source, (NH4)2HPO4 as a nitrogen and phosphorus source, KH2PO4 as a buffer agent, and Mg2SO4 · 7H2O as a magnesium and sulfur source were used as variables in the medium. The optimum medium composition was found to be 34.694 g L?1 of reducing sugar (equivalent to molasses solution), 8.536 g L?1 of (NH4)2HPO4, 3.073 g L?1 of KH2PO4, and 1.707 g L?1 of Mg2SO4 · 7H2O. Molasses-based culture medium increased the yield of cmFDH about three times compared to LB medium. The currently developed media has the potential to be used in industrial bioprocesses with low-cost production.  相似文献   

5.
In pH 6.0 Na2HPO4-NaH2PO4 buffer solution and in the presence of cetyltrimethyl ammonium bromide, nanosilver particles were aggregated to a stable suspension. Therein, rhodamine 6G (Rh6G) exhibited three strong surface-enhanced Raman scattering (SERS) peaks at 613, 1,363, and 1,510 cm?1, and their SERS intensities were enhanced when the concentration of Rh6G increased. In the presence of Hg2+, the SERS intensity decreased greatly owing to formation of stable Rh6G-HgBr 4 2? ternary association complex molecules as well as its particles. In the optimal condition, the decreased SERS intensity at 613 cm?1 responds linearly with the concentration of Hg2+ over 25–2,000 nmol/L. Thus, a new sensitive SERS method has been proposed for the determination of trace Hg2+ in the water sample, with satisfactory results.  相似文献   

6.
Residue-specific amide proton spin-flip rates K were measured for peptide-free and peptide-bound calmodulin. K approximates the sum of NOE build-up rates between the amide proton and all other protons. This work outlines the theory of multi-proton relaxation, cross relaxation and cross correlation, and how to approximate it with a simple model based on a variable number of equidistant protons. This model is used to extract the sums of K-rates from the experimental data. Error in K is estimated using bootstrap methodology. We define a parameter Q as the ratio of experimental K-rates to theoretical K-rates, where the theoretical K-rates are computed from atomic coordinates. Q is 1 in the case of no local motion, but decreases to values as low as 0.5 with increasing domination of sidechain protons of the same residue to the amide proton flips. This establishes Q as a monotonous measure of local dynamics of the proton network surrounding the amide protons. The method is applied to the study of proton dynamics in Ca2+-saturated calmodulin, both free in solution and bound to smMLCK peptide. The mean Q is 0.81 ± 0.02 for free calmodulin and 0.88 ± 0.02 for peptide-bound calmodulin. This novel methodology thus reveals the presence of significant interproton disorder in this protein, while the increase in Q indicates rigidification of the proton network upon peptide binding, confirming the known high entropic cost of this process.  相似文献   

7.
We show that DNA carrying 5-methylcytosine modifications or methylated DNA (m-DNA) can be distinguished from DNA with unmodified cytosine by Raman spectroscopy enhanced by both a bowtie nanoantenna and excitation resonance. In particular, m-DNA can be identified by a peak near 1000 cm?1 and changes in the Raman peaks in the 1200–1700 cm?1 band that are enhanced by the ring-absorption resonance. The identification is robust to the use of resonance Raman and nanoantenna excitation used to obtain significant signal improvement. The primary differences are three additional Raman peaks with methylation at 1014, 1239, and 1639 cm?1 and spectral intensity inversion at 1324 (C5=C6) and 1473 cm?1 (C4=N3) in m-DNA compared to that of DNA with unmodified cytosine. We attribute this to the proximity of the methyl group to the antenna, which brings the (C5=C6) mode closer to experiencing a stronger near-field enhancement. We also show distinct Raman spectral features attributed to the transition of DNA from a hydrated state, when dissolved, to a dried/denatured state. We observe a general broadening of the larger lines and a transfer of spectral weight from the ~1470 cm?1 vibration to the two higher-energy lines of the dried m-DNA solution. We attribute the new spectral characteristics to DNA softening under high salt conditions and find that the m-DNA is still distinguishable via the ~1000 cm?1 peak and distribution of the signal in the 1200–1700 cm?1 band. The nanoantenna gain exceeds 20,000, whereas the real signal ratio is much less because of a low average enhanced region occupancy even with these relatively high DNA concentrations. It is improved when fixed DNA in a salt crystal lies near the nanoantenna. The Raman resonance gain profile is consistent with A-term expectations, and the resonance is found at ~259 nm excitation wavelength.  相似文献   

8.
Organic soiling is a major issue in the food processing industries, causing a range of biofouling and microbiological problems. Energy dispersive X-ray (EDX) and Fourier transform infra red spectroscopy (FT-IR) were used to quantify and determine the biochemical groups of food soils on stainless steel surfaces. EDX quantified organic material on surfaces where oily based residues predominated, but was limited in its usefulness since other food soils were difficult to detect. FT-IR provided spectral ‘fingerprints’ for each of the soils tested. Key soiling components were associated with specific peaks, viz. oils at 3025 cm?1–3011 cm?1, proteins at 1698 cm?1–1636 cm?1 and carbohydrates at 1658 cm?1–1596 cm?1, 783 cm?1–742 cm?1. High concentrations of some soils (10%) were needed for detection by both EDX and FT-IR. The two techniques may be of use for quantifying and identifying specific recalcitrant soils on surfaces to improve cleaning and hygiene regimes.  相似文献   

9.
Coumarin molecules have biological activities possessing lipid-controlling activity, anti-hepatitis C activity, anti-diabetic, anti-Parkinson activity, and anti-cancer activity. Here, we have presented an inclusive study on the interaction of 8-substituted-7-hydroxy coumarin derivatives (Umb-1/Umb-2) with α-1-glycoprotein (AGP) and human serum albumin (HSA) which are the major carrier proteins in the human blood plasma. Binding constants obtained from fluorescence emission data were found to be KUmb-1=3.1 ± .01 × 104 M?1, KUmb-2 = 7 ± .01 × 104 M?1, which corresponds to ?6.1 and ?6.5 kcal/mol of free energy for Umb-1 and Umb-2, respectively, suggesting that these derivatives bind strongly to HSA. Also these molecules bind to AGP with binding constants of KUmb-1-AGP=3.1 ± .01 × 103 M?1 and KUmb-2-AGP = 4.6 ± .01 × 103 M?1. Further, the distance, r between the donor (HSA) and acceptor (Umb-1/Umb-2) was calculated based on the Forster’s theory of non-radiation energy transfer and the values were observed to be 1.14 and 1.29 nm in Umb-1–HSA and Umb-2–HSA system, respectively. The protein secondary structure of HSA was partially unfolded upon binding of Umb-1 and Umb-2. Furthermore, site displacement experiments with lidocaine, phenylbutazone (IIA), and ibuprofen (IIIA) proves that Umb derivatives significantly bind to subdomain IIIA of HSA which is further supported by docking studies. Furthermore, Umb-1 binds to LYS402 with one hydrogen bond distance of 2.8 Å and Umb-2 binds to GLU354 with one hydrogen bond at a distance of 2.0 Å. Moreover, these molecules are stabilized by hydrophobic interactions and hydrogen bond between the hydroxyl groups of carbon-3 of coumarin derivatives.  相似文献   

10.
In the coupling of ATP pyrophosphorolysis to Ca2+ transport in beef heart mitochondria, for each molecule of ATP cleaved, one proton is released and one Ca2+ is transported into the interior space. With the use of tritium labelled ATP, it could be shown that ATP is pyrophosphorylyzed into a species equivalent to Pi that moves inward, and a species equivalent to ADP that is extruded into the aqueous space on the exterior of the mitochondrion. The species equivalent to Pi has been identified as a negatively charged form of Pi (PO?) and the species equivalent to ADP as a positively charged form (ADP+). The inward flow of PO? is coupled to the inward flow of Ca2+ in 1:1 stoichiometry—a token that Ca2+ must enter in the form of Ca2+A?, where A? is a monovalent anion. During ATP pyrophosphorolysis protons are released on the I side and taken up on the M side—one proton for each molecule of ATP cleaved. The alkalinization of the matrix space leads to the deposition of Ca3(PO4)2 and to the extrusion of the two species released by this deposition (Pi, K+). Two thirds of the PO? is trapped as Ca3(PO4)2 in the matrix space and one third is extruded into the external space. The extrusion of K+ provides a mechanism by which protons can be continuously brought into the matrix space to sustain a high rate of coupled pyrophosphorolysis of ATP. The coupling pattern for Ca2+ transport driven by ATP pyrophosphorolysis is identical with the corresponding pattern for Ca2+ transport driven by electron transfer. This identity is suggestive that coupling mediated by the Fo-F1 system and coupling mediated by electron transfer complexes are mechanistically indistinguishable.  相似文献   

11.
High salinity wastewaters have limited treatment options due to the occurrence of salt inhibition in conventional biological treatments. Using recirculating marine aquaculture effluents as a case study, this work explored the use of Constructed Wetlands as a treatment option for nutrient and salt loads reduction. Three different substrates were tested for nutrient adsorption, of which expanded clay performed better. This substrate adsorbed 0.31 mg kg?1 of NH4 +?N and 5.60 mg kg?1 of PO4 3??P and 6.9 mg kg?1 dissolved salts after 7 days of contact. Microcosms with Typha latifolia planted in expanded clay and irrigated with aquaculture wastewater (salinity 2.4%, 7 days hydraulic retention time, for 4 weeks), were able to remove 94% NH4 +?N (inlet 0.25 ± 0.13 mg L?1), 78% NO2 ??N (inlet 0.78 ± 0.62 mg L?1), 46% NO3 ??N (inlet 18.83 ± 8.93 mg L?1) whereas PO4 3??P was not detected (inlet 1.41 ± 0.21 mg L?1). Maximum salinity reductions of 52% were observed. Despite some growth inhibition, plants remained viable, with 94% survival rate. Daily treatment dynamics studies revealed rapid PO4 3??P adsorption, unbalancing the N:P ratio and possibly affecting plant development. An integrated treatment approach, coupled with biomass valorization, is suggested to provide optimal resource management possibilities.  相似文献   

12.
13.
A set of supramolecular cage-structures—spherophanes—was studied at the density functional B3LYP level. Full geometrical structure optimisations were made with 6–31G and 6–31G(d) basis sets followed by frequency calculations, and electronic energies were evaluated at B3LYP/6–31++G(d,p). Three different symmetries were considered: C1, Ci, and Oh. It was found that the bonds between the benzene rings are very long to allow π-electron delocalisation between them. These spherophanes show portal openings of 2.596 Å in Spher1, 4.000 Å in Meth2, 3.659 Å in Oxa3, and 4.412 Å in Thia4. From the point of view of potential host–guest interaction studies, it should also be noted that the atoms nearest to the centre of the cavities are carbons bonded to X groups. These supramolecules seem to exhibit relatively large gap HOMO?LUMO: 2.89 eV(Spher1), 5.26 eV(Meth2), 5.73 eV(Oxa3), and 4.82 eV(Thia4). The calculated ΔH°f (298.15 K) values at B3LYP/6–31G(d) are (in kcal mol?1) 750.98, 229.78, ?10.97, and 482.49 for Spher1, Meth2, Oxa3, and Thia4, respectively. Using homodesmotic reactions, relative to Spher1, the spherophanes Meth2, Oxa3, and Thia4 are less strained by ?399.13 kcal mol?1, ?390.40 kcal mol?1, and ?411.38 kcal mol?1, respectively. Their infrared and 13C NMR calculated spectra are reported.  相似文献   

14.
Recent measurements have demonstrated unprecedented increase in atmospheric deposition of nutrients in many parts of India. To determine whether atmospheric nutrient inputs would increase phytoplankton growth and catchment dissolved organic carbon (DOC) flushing to constrain benthic algae, we analyzed NO3 ? and PO 4 ?3 in atmospheric deposits; nutrients and DOC in runoff and lake water and standing crop biomass of phytoplankton and periphyton at Jaisamand Lake of Rajasthan, India. Atmospheric deposition of NO3 ? (7.18–29.95 kg ha?1 year?1) and PO 4 ?3 (0.56–2.15 kg ha?1 year?1) showed a consistently rising trend across the year. Microbial biomass and activity in catchment increased in response to atmospheric deposition. Lake DOC and nutrients showed strong coherence with their terrestrial and atmospheric fluxes. Phytoplankton development showed significant linearity with atmospheric input of nutrients. Air-driven input appeared to have compensated the nutrient constraints to phytoplankton during drought. The N:P stoichiometry of deposition and that of lake water indicated that, although there was a seasonal switchover to N- or P-limitation, phytoplankton were mainly co-limited by N and P due probably to the synergistic effects of combined N + P enrichment in the pelagic zone of the lake. Periphyton standing crop showed inverse relationship with phytoplankton and lake DOC. The study indicated that enhanced phytoplankton development and terrestrial DOC flushing in response to atmospheric nutrient input attenuated light penetration to constrain algal periphyton. We suggests that data on these issues may be considered in developing aquatic ecosystem models to establish future links between changing air–water–land interactions and associated shifts in lake ecosystem functioning for more accurately predicting climate change drivers and designing integrated lake basin management strategies.  相似文献   

15.
Raman studies of conformational changes in model membrane systems   总被引:2,自引:0,他引:2  
Laser Raman spectra of concentrated samples of phosphatidyl choline and phosphatidyl ethanolamine were taken at approximately 10° intervals over a temperature range of 90°–19°C. The spectral region from 30 to 3300 cm?1 was investigated. Several new spectral features were discovered which are correlated to phospholipid liquid crystalline structure. It is shown that 1) frequency shifts occur in the PO2? symmetric stretch band which suggest a change in exposure of the PO2 group to the solvent upon melting, 2) the frequency of the translational hydrocarbon mode around 150 cm?1 appears to indicate the degree to which the hydrocarbon chain is extended, 3) the methyl and methylene stretch bands at 2890 and 2850 cm?1 very clearly demonstrate hydrocarbon chain melting, and 4) the 720 cm?1 band, previously assigned to the symmetric OPO diester stretch, appears to be due instead to the symmetric CN stretch of choline.  相似文献   

16.
A theoretical study of the magnetic properties, using density functional theory, of a family of trinuclear μ3-OH copper(II) complexes reported in the literature is presented. The reported X-ray crystal structures of [Cu33-OH)(aat)3(H2O)3](NO3)2·H2O (HUKDUM), where aat: 3-acetylamine-1,2,4-triazole; [Cu33-OH)(aaat)3(H2SO4)(HSO4)(H2O)] (HUKDOG), where aaat: 3-acetylamine-5-amine-1,2,4-triazole; [Cu33-OH)(PhPyCNO)3(tchlphac)2] (HOHQUR), where PhPyCNO: phenyl 2-pyridyl-ketoxime and tchlphac: acid 2,4,5-trichlorophenoxyacetic; [Cu33-OH)(PhPyCNO)3(NO3)2(CH3OH)] (ILEGEM); [Cu33-OH)(pz)3(Hpz)3(ClO4)2] (QOPJIP), where Hpz?=?pyrazole; [Cu33-OH)(pz)3(Hpz)(Me3CCOO)2]?2Me3CCOOH (DEFSEN) and [Cu33-OH)(8-amino-4-methyl-5-azaoct-3-en-2-one)3][CuI3] (RITXUO), were used in the calculations. The magnetic exchange constants were calculated using the broken-symmetry approach. The calculated J values are for HUKDUM J1?=??68.6 cm?1, J2?=??69.9 cm?1, J3?=??70.4 cm?1; for HUKDOG, J1?=??73.5 cm?1, J2?=??58.9 cm?1, J3?=??62.1 cm?1; for HOHQUR J1?=??128.3 cm?1, J2?=??134.1 cm?1, J3?=??120.4 cm?1; for ILEGEM J1?=??151.6 cm?1, J2?=??173.9 cm?1, J3?=??186.9 cm?1; for QOPJIP J1?=??118.3 cm?1, J2?=??106.0 cm?1, J3?=??120.6 cm?1; for DEFSEN J1?=??74.9 cm?1, J2?=??64.0 cm?1, J3?=??57.7 cm?1 and for RITXUO J1?=??10.9 cm?1, J2?=?+14.3 cm?1, J3?=??35.4 cm?1. The Kahn-Briat model was used to correlate the calculated magnetic properties with the overlap of the magnetic orbitals. Spin density surfaces show that the delocalization mechanism is predominant in all the studied compounds.
Figure
The Kahn-briat model was used to correlate the calculated magnetic properties with the overlap of the magnetic orbitals.  相似文献   

17.
The environmental benefits of fuel cells and electrolyzers have become increasingly recognized in recent years. Fuel cells and electrolyzers that can operate at intermediate temperatures (300–450 °C) require, in principle, neither the precious metal catalysts that are typically used in polymer‐electrolyte‐membrane systems nor the costly heat‐resistant alloys used in balance‐of‐plant components of high‐temperature solid oxide electrochemical cells. These devices require an electrolyte with high ionic conductivity, typically more than 0.01 S cm?1, and high chemical stability. To date, however, high ionic conductivities have been found in chemically unstable materials such as CsH2PO4, In‐doped SnP2O7, BaH2, and LaH3?2xOx. Here, fast and stable proton conduction in 60‐at% Sc‐doped barium zirconate polycrystal, with a total conductivity of 0.01 S cm?1 at 396 °C for 200 h is demonstrated. Heavy doping of Sc in barium zirconate simultaneously enhances the proton concentration, bulk proton diffusivity, specific grain boundary conductivity, and grain growth. An accelerated stability test under a highly concentrated and humidified CO2 stream using in situ X‐ray diffraction shows that the perovskite phase is stable over 240 h at 400 °C under 0.98 atm of CO2. These results show great promises as an electrolyte in solid‐state electrochemical devices operated at intermediate temperatures.  相似文献   

18.
We investigated vacancy-assisted self-diffusion in germanium by means of kinetic lattice Monte Carlo (KLMC) simulations below the melting temperature, for a vacancy concentration of 1 × 1018/cm3. At higher temperatures, fewer clusters formed, but there was less variation in the number of clusters than at lower temperatures as the time increased. Equilibrium diffusivities in the clustering region were 102 lower than those of free vacancies in the initial stage of KLMC simulations. They were expressed according to three temperature regimes: 6.5 × 10? 4 exp(–0.35/k B T) cm2/s at temperatures above 1100 K, 5.2 × 105 exp(–2.32/k B T) cm2/s at temperatures of 900–1100 K and 6.0 × 0–7 exp(–0.19/k B T) cm2/s at temperatures below 900 K. The effective mean migration energy, 1.1 eV, closely coincided with that of the 1.0–1.2 eV in experiments and was very different from the migration energy of the free vacancy.  相似文献   

19.
The structure, electron density distribution, energetic and electrostatic properties of simple nitramine based energetic TMA, DMNA, MDA and TNA molecules were determined using density functional theory (B3LYP) with the 6-311G** and aug-cc-pVDZ basis sets coupled with Bader's theory of atoms in molecules. In the NO2 group substituted molecules, the N–N bond distance increases with the increase of NO2 groups, whereas in C–N bonds, this effect is relatively less, and the distances are almost equal. The topological analysis of electron density reveals that the electron density ρbcp(r) of C–N and N–N bonds are significantly decreasing with the increase of NO2 groups in the nitramine molecules. The Laplacian of electron density ▽2ρbcp(r) of N–NO2 bonds [DMNA: ? 16.7 eÅ? 5, MDA: ? 12.8 eÅ? 5 and TNA: ? 7.9 eÅ? 5] of the molecules are relatively less negative, and the values also decrease with the increase of NO2 groups; this implies that the charge concentration decreases with the increase of NO2 groups, which leads to weakening the N–N bonds of the molecules. The isosurface of molecular electrostatic potential displays high electronegative regions around the NO2 groups. The oxygen balance OB100 of the molecules increases as the number of NO2 group increases in the molecules, in which, the TNA molecule having maximum OB100 value [+7.89]. The band gap, heat of detonation, bond dissociation energy and charge imbalance are predominantly depends on the number of NO2 group present in the molecule. The charge imbalance parameter (ν) has been calculated for all molecules, which reveals that TNA is a highly sensitive molecule, the corresponding ν value is 0.047.  相似文献   

20.
Computer simulation method was applied to investigate the migration of lithium ion in three amorphous solid systems containing polyoxovanadate (POV) clusters [V10O28]6 ? . The cluster was adopted from a recently synthesized crystalline poly[octa-μ-aqua-octaaqua-μ-decavanadato-hexalithium] (POAODH). The simulated POV systems correspond to amorphous solid half-dehydrated solid and completely dehydrated solid doped with LiCl salt. The simulation results show large diffusion constants of lithium ions in all systems in spite of highly negatively charged [V10O28]6 ?  clusters presented in the system. The estimated ionic conductivity due to the migration of lithium ions reaches a magnitude of 10? 4 S/m. The conductivity increases as the water content in the system decreases. The analysis of moving trajectories shows the lithium ion moves around the oxygen sites of POV clusters and hops between them. The estimated displacement of lithium ion is about 4~5 Å, which is much larger than the corresponding displacement of lithium ion in a polymer matrix. Rapidly rotating clusters shown by orientation correlation function analysis, in conjunction with the large separation between clusters in the system, provides favorable conditions for the large amplitude migration of lithium ions.  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