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 125 毫秒
1.
We performed a series of molecular dynamics (MD) simulations on Nafion® membranes containing various quantities of H2O and CH3OH. The simulations afforded diverse nanoscale phase-separated structures, such as clusters, channels, and cluster–channels. The calculated cluster–channel structure qualitatively agrees with the experimental results of X-ray diffraction studies. We also investigated the diffusion mechanisms for H2O, protons, CH3OH, H2, and O2 in these membranes. To reproduce the hopping transfer of protons, we employed a semi-classical MD approach using the empirical valence bond method. The estimated diffusion coefficients of H2O and proton in the membranes significantly depended on the H2O content, and these values showed qualitatively good agreement with the experimental results. The diffusion coefficient of proton in H2O-rich membranes was much larger than that of H2O, and the proton mainly formed H5O2+ complex. Furthermore, the simulation results indicate that the majority of CH3OH permeates through the H2O clusters, and the majority of H2 and O2 permeates through the hydrophobic region of the membrane.  相似文献   

2.
The energy interaction curves of a number of diatomic and polyatomic dication systems were calculated in order to study their energy-trapping properties. Generally, the ab initio complete active space multiconfiguration self-consistent field method was used in an extended valence + polarization basis set, with compact effective potentials replacing the core electrons. The diatomic dications include all ten possible binary combinations of oxygen, sulphur, selenium, and tellurium. O22+ shows the largest exothermicity, measured from equilibrium to the monocation combination asymptote, and highest barrier to dissociation. The calculated equilibrium bond length and harmonic vibrational frequency agree very well with experiment. The O22+, SO2+, SeO2+, and TeO2+ series show progressively decreasing exothermicities but similar barrier heights. The non-oxides, in contrast, show similar exothermicities but decreasing barriers with increasing size of the atom constituents. These trends are interpreted in terms of both valence bond curve-crossing and molecular orbital bonding models. The ozone dication, O32+, is found to have a number of low-lying singlet and triplet stationary state structures spanning near-linear to D3h2+ symmetries. Although the calculated exothermicity is even larger than for O22+, the barrier to O2+ + O+ dissociation is predicted to be low in each case. O22+ surrounded by six argon atoms to model an isolating environment shows increased equilibrium O–O bond length, decreased exothermicity, and increased barrier to dissociation, relative to the bare dication. O22+ flanked at each end by a perpendicularly oriented H2 molecule in a staggered conformation is obstructed from direct conversion to the water dimer dication by a high barrier. However, [(H2O)2]2+ dissociates smoothly from equilibrium to two water monocations with a large exothermicity but a small barrier.  相似文献   

3.
《Ceramics International》2016,42(16):18373-18379
This paper reports the performance of an yttria-stabilized zirconia fuel cell (YSZ) using five kinds of gas systems. The final target of this research is to establish the combined fuel cell systems which can produce a H2 fuel and circulate CO2 gas in the production process of electric power. A large electric power was measured in the H2–O2 gas system and the CO–O2 gas system at 1073 K. The formation process of O2− ions in the endothermic cathodic reaction (1/2O2+2e→O2−) controlled the cell performance in both the gas systems. The electric power of the H2–CO2 gas system, which allowed to change CO2 gas into a CO fuel (H2+CO2→H2O+CO) in the cathode, was 1/31–1/11 of the maximum electric power for the H2–O2 gas system. This result is related to the larger endothermic energy for the formation of O2− ions from CO2 molecules at the cathode (CO2+2e→CO+O2−) than from O2 molecules. The CO–H2O gas system and the H2–H2O gas system was expected to produce a H2 fuel in the cathode (CO+H2O→H2+CO2, H2+H2O→H2+H2O). Although relatively high OCV values (open circuit voltage) were measured in these gas systems, no electric power was measured. At this moment, it was difficult to apply H2O vapor as an oxidant to the cathodic reaction in a YSZ fuel cell.  相似文献   

4.
The Se(VI)-analogues of ettringite and monosulfate, selenate-AFt (3CaO·Al2O3·3CaSeO4·37.5H2O), and selenate-AFm (3CaO·Al2O3·CaSeO4·xH2O) were synthesised and characterised by bulk chemical analysis and X-ray diffraction. Their solubility products were determined from a series of batch and resuspension experiments conducted at 25 °C. For selenate-AFt suspensions, the pH varied between 11.37 and 11.61, and a solubility product, log Kso=61.29±0.60 (I=0 M), was determined for the reaction 3CaO·Al2O3·3CaSeO4·37.5H2O+12 H+⇔6Ca2++2Al3++3SeO42−+43.5H2O. Selenate-AFm synthesis resulted in the uptake of Na, which was leached during equilibration and resuspension. For the pH range of 11.75 to 11.90, a solubility product, log Kso=73.40±0.22 (I=0 M), was determined for the reaction 3CaO·Al2O3·CaSeO4·xH2O+12 H+⇔4Ca2++2Al3++SeO42−+(x+6)H2O. Thermodynamic modelling suggested that both selenate-AFt and selenate-AFm are stable in the cementitious matrix; and that in a cement limited in sulfate, selenate concentration may be limited by selenate-AFm to below the millimolar range above pH 12.  相似文献   

5.
The influence of proton hydration in Wells-Dawson H6P2W18O62 (HP2W) and Keggin H3PW12O40 (HPW) type heteropolyacids on their catalytic activity in ethyl-tert-butyl ether (ETBE) synthesis in gas phase was investigated. In the case of samples with the content of crystallisation water per one proton below 0.8 ([H2O]/[H+] < 0.8), the catalytic activity related to the mass unit of the HP2W was higher than that of the HPW in accordance with the order of proton concentration in mass unit. On the other hand, the activity related to one mole of protons indicated higher activity in the HPW than in the HP2W in accordance with the order of acid strength. In the series of samples with [H2O]/[H+] > 0.8 the activity was much higher for the HP2W than the HPW and maximum of both catalytic activity and ethanol sorption capacity was observed for the HP2W. The possible role of the secondary structure of the hydrates was discussed.  相似文献   

6.
Five protonic beta and beta″-aluminas; viz hydrated sodium beta-alumina (1,24Na2O·11Al2O3), hydronium beta-alumina (1.24H2O·11Al2O3·2.6H2O), partially dehydrated hydronium beta-alumina (1.24H2O·11Al2O3·1.3H2O), hydrogen beta-alumina (1.24H2O·11Al2O3) and hydronium beta″-alumina (0.84H2O·0.8MgO·5Al2O3·2.8H2O) were examined by broad band nuclear magnetic resonance from ?196°C to 200°C. The spectra of hydronium beta-alumina and hydronium beta″-alumina are consistent with a mixed composition of H2O, H3O+ and H+ species in the conducting plane. Hydrogen beta-alumina and partially dehydrated hydronium beta-alumina appear to contain only relatively isolated (2.6–2.7Å) protons; no evidence of molecular water or hydronium ions is found. Water molecules intercalated into the conduction plane of sodium beta-alumina do not appear to be in rapid motion, even at 167°C, but are relatively stationary. The onset of motional narrowing in hydronium beta″-alumina occurs at ?40°C but not until +30°C in hydronium beta-alumina. This is consistent with the higher conductivity reported for hydronium beta″-alumina, 10?3–10?5 (ohm-cm)?1 at 25°C, in comparison to 10?10–10?11 (ohm-cm)?1 for hydronium beta-alumina at 25°C.  相似文献   

7.
The classical and quantum equilibrium properties of an excess proton in bulk phase water are examined computationally with a special emphasis on the influence of an explicit quantum dynamical treatment of the nuclei on the calculated observables. The potential model used, our recently developed multistate empirical valence bond (MS-EVB) approach is described. The MS-EVB model takes into account the interaction of an exchange charge distribution of the charge-transfer complex with the polar solvent, which qualitatively changes the nature of the solvated complex. The impact and importance of the exchange term on the stability of the solvated H5O2+ (Zundel) cation relative to the H9O4+ (Eigen) cation in the liquid phase is demonstrated. Classical and quantum path-integral molecular dynamics (PIMD) simulations of an excess proton in bulk phase water reveal that quantization of the nuclear degrees of freedom results in an increased stabilization of the solvated Zundel cation relative to the Eigen cation, and that species intermediate between the two are also probable. Quantum effects lead to a significant broadening of the probability distributions used to characterize the two species, and a definite differentiation and sharp characterization of the species connected to the excess proton in liquid water is found to be difficult.  相似文献   

8.
A new inorganic–organic hybrid material (C10H8N2H)2(Mo8O26) · 2NH4 (1) was prepared under solvothermal conditions (H2O/ethanol), which was then used as a starting material in the synthesis of another material, (C10H8N2Cu)2(Mo8O26) · 2NH4 (2). Each compound has been characterized by single-crystal X-ray diffraction, X-ray powder diffraction, elemental analysis, thermogravimetric analysis, fluorescence, FT-IR, and bond valence sum calculations. The comparative study focuses on the transformation and structural change between 1 and 2, which arises from the simple substitution of the proton in 1 with the Cu+ ions. The other surprising feature is that Cu+ ions can not only directly be used to substitute the proton of 1 but also ameliorate its photoluminescence property. It means that 1 appears to have recognizable functional protons in the solid-state structure and, hence, can very well be considered as an ideal precursor agent to design the functional materials with special architectures and photoactive properties.  相似文献   

9.
Experimental evidence for proton solvation and proton mobility is analyzed and the results are compared with recent simulations. Three factors contribute to differences in proton solvation energies: hydrogen-bond cleavage, changes in hydrogen-bond lengths, and proton derealization. These factors are estimated from experimental attributes. In dilute acidic aqueous solutions H3O+ is more stable than H5O2+ by about 0.6 kcal/mol. This estimate, together with the activation energy for proton mobility, supports the 121 mechanism for proton mobility in which a protonated water monomer is transformed, by second-shell hydrogen-bond cleavage, to a protonated dimer and back to another protonated monomer.  相似文献   

10.
Ozone reacts slowly with Ag+ (circumneutral pH, k = (11 ± 3) × 10?2 M?1 s?1). After some time, ozone decay kinetics may suddenly become faster with the concomitant formation of silver sol. As primary process, an O-transfer from O3 to Ag(I) is suggested, whereby Ag(III) is formed [Ag+ + O3 + 2 H2O → Ag(OH)3 + O2 + H+]. This conproportionates with Ag(I), which is in large excess, leading to Ag(II) [Ag+ + Ag(OH)3 ? 2 Ag(OH)+ + HO?]. Further, Ag(II) reacts with ozone in a high exergonic reaction [Ag(OH)+ + O3 → Ag + 2 O2 + H+], where ozone acts as a reducing agent. Thereby, a single silver atom, Ag, is formed that can be oxidized by O2 and O3 or can aggregate to a silver sol. Aggregation slows down the rate of oxidation. When Ag+ is complexed by acetate ions, ozone decay and silver sol formation are speeded up by enhancing Ag(II) formation [Ag(I)acetate + O3 → Ag(III)acetate → Ag(II) + CO2 + ?CH3]. In the presence of oxalate, the formed complex reacts faster with ozone than Ag+, and Ag(III)oxalate decarboxylates rapidly [Ag(I)oxalate + O3 → Ag(III)oxalate → Ag+ + 2 CO2]. This enhances ozone decay but prevents silver sol formation. Quantum chemical calculations have been carried out for substantiating mechanistic suggestions.  相似文献   

11.
Graphite paste electrode allows to determine elementary processes of the electrochemical oxidation in aqueous media of an electrochemical probe such as: N-acetyl L-tyrosine amide. Mathematical analysis of voltammograms gives the following EC mechanism: R?C6H5OH?R?C6H5O. + H+ + e 2 R?C6H5O.R?C6H5O+ + R?C6H5O?, R?C6H5O? + H+R?C6H5OH, R?C6H5O+ → [R?C6H4O].. + H+, n[R?C6H4O].. → ?[R?C6H4O]?n.  相似文献   

12.
Bottleneck size is the minimum Li+ migration channel of Li7La3Zr2O12 (LLZO) and it greatly influences the Li+ conductivity. Doping different elements on the Zr site of LLZO can adjust the bottleneck size and improve the Li+ conductivity. However, the regulation mechanism is not clear. In this work, Li6.4La3Zr1.4M0.6O12 (M = Sb, Ta, Nb) has been prepared and the bottleneck size has been adjusted by doping different pentavalent ions. The results manifest that the cell parameter and bottleneck size decrease with the rise of the radius of doped pentavalent ions. This is because larger pentavalent ion leads to larger bond length of M–O, and weaker covalent component between M5+ and O2-, corresponding, the formal charge on the M5+ become larger, and the bond length of La–O slightly decreases due to the coulomb repulsion between La3+ and M5+ increase. While, the activation energy drop firstly and then rise with the rise of bottleneck size because of the migration of Li+ not only relate to the size of the migration channel but also to the strength of M–O covalent bonding. The bottleneck size and bond length of M–O synergistically affect the migration of Li+.  相似文献   

13.
The electrochemistry of molten LiOH–NaOH, LiOH–KOH, and NaOH–KOH was investigated using platinum, palladium, nickel, silver, aluminum and other electrodes. The fast kinetics of the Ag+/Ag electrode reaction suggests its use as a reference electrode in molten hydroxides. The key equilibrium reaction in each of these melts is 2 OH = H2O + O2– where H2O is the Lux-Flood acid (oxide ion acceptor) and O2– is the Lux–Flood base. This reaction dictates the minimum H2O content attainable in the melt. Extensive heating at 500 °C simply converts more of the alkali metal hydroxide into the corresponding oxide, that is, Li2O, Na2O or K2O. Thermodynamic calculations suggest that Li2O acts as a Lux–Flood acid in molten NaOH–KOH via the dissolution reaction Li2O(s) + 2 OH = 2 LiO + H2O whereas Na2O acts as a Lux–Flood base, Na2O(s) = 2 Na+ + O2–. The dominant limiting anodic reaction on platinum in all three melts is the oxidation of OH to yield oxygen, that is 2 OH 1/2 O2 + H2O + 2 e. The limiting cathodic reaction in these melts is the reduction of water in acidic melts ([H2O] [O2–]) and the reduction of Na+ or K+ in basic melts. The direct reduction of OH to hydrogen and O2– is thermodynamically impossible in molten hydroxides. The electrostability window for thermal battery applications in molten hydroxides at 250–300 °C is 1.5 V in acidic melts and 2.5 V in basic melts. The use of aluminum substrates could possibly extend this window to 3 V or higher. Preliminary tests of the Li–Fe (LAN) anode in molten LiOH–KOH and NaOH–KOH show that this anode is not stable in these melts at acidic conditions. The presence of superoxide ions in these acidic melts likely contributes to this instability of lithium anodes. Thermal battery development using molten hydroxides will likely require less active anode materials such as Li–Al alloys or the use of more basic melts. It is well established that sodium metal is both soluble and stable in basic NaOH–KOH melts and has been used as a reference electrode for this system.  相似文献   

14.
A novel hexagonal-based honeycomb compound with overall formula {[KCr(C2O4)3][Cu(pypn)(H2O)](H2O)4} is reported in which pypn is with the tetradentate ligand (N,N′-bis(2-pyridylmethyl)-1,3-propanediamine). The [KCr(C2O4)3]2? moiety forms a hexagonal honeycomb structure, while the five-coordinated [Cu(pypn)(H2O)]2+ moiety is located in between the layers, partly filling the holes in the cavities. The synthesis, X-ray crystal structure and some spectroscopic properties are presented. The coordination of Cr(III) is octahedral, with a CrO6 chromophore, and the K+ ion is in a KO6 environment (K–O distances vary from 2.36 to 2.48 ?). The [KCr(C2O4)3]2? layers have the K+ ions in a Λ conformation, while the Cr(III) ions in the Δ conformation. The geometry around the Cu(II) is five-coordinated with four nitrogens from the chelating pypn ligand in a plane and the apical position being occupied by the oxygen atom of the coordinating water molecule. The packing of the cationic and the anionic layers appears to be of special interest.  相似文献   

15.
Electrode-potential-dependent activation energies for electron transfer have been calculated using a local reaction center model and constrained variation theory for the oxygen reduction reaction on platinum in base. Results for four one-electron transfer steps are presented. For the first, O2(ads) is predicted to be reduced to adsorbed superoxide, O2(ads), which dissociates with a low activation barrier to O(ads) + O(ads). Then a proton transfer form H2O(ads) to O(ads) takes place, forming OH(ads) + OH(aq). The second electron transfer reacts O(ads) with H2O(aq) to form a second OH(ads) + OH(aq). The third and fourth electron transfers react the two OH(ads) with two H2O(aq) to form two H2O(ads) + two OH(aq). All three different surface reduction reactions are predicted to have reversible potentials in the −0.24 V(SHE) to −0.29 V(SHE) range for 0.1 M base and activation energies for the superoxide formation step are close to the experimentally observed range in 0.1 M base for the overall four-electron to water over the three low index (1 1 0) (1 0 0) and (1 1 1) surfaces: 0.38-0.49 eV at 0.35 eV respectively at 0.88 V(RHE). Predicted reversible potentials for forming O2(ads) are compared with estimates from the experimental literature. The difference between the acid mechanism, where the peroxyl radical, OOH(ads) is the first reduction intermediate, and the base mechanism, where superoxide, O2(ads) is the first reduction intermediate, is discussed.  相似文献   

16.
Dibarium silicate, Ba2SiO4, was hydrated in two ways: in paste form at 25° using a water/solid weight ratio of 0.7:1, and in a polyethylene bottle rotated on a wheel at 5°, 25° and 50°, using a water/solid weight ratio of 9.0:1. When Ba2SiO4 is hydrated in paste form, the stoicheiometry of the reaction at 25° is the same as in bottle-hydration at 50°: 2BaO.SiO2+2.2H2O = 1.2BaO.SiO2.1.4H2O+0.8Ba(OH)2. The stoicheiometry of bottle-hydration at 5° and 25° is represented by the equation: 2BaO.SiO2+2.2H2O = BaO.SiO.1.2H2O+Ba(OH)2. Barium silicate hydrate, 1.2BaO.SiO2.1.4H2O, is well crystallised and has a specific surface area of ? 3m2/g. The crystals are plate-like and have a tendency to form clusters. The low-baria hydrate, BaO.SiO2.1.2H2O, is poorly crystallised and consists of thin platelets. It has a specific surface area of ? 35m2/g. The thermal dehydration of fully hydrated barium silicate and of the barium silicate hydrates was investigated by thermogravimetric and differential thermal analysis techniques. The similarities and differences between the barium silicate hydrates obtained in the hydration of barium silicate and the calcium silicate hydrates obtained in the hydration of β-dicalcium silicate and Ca3SiO5 are discussed. A mechanism of hydration of barium silicate is proposed which involves solution, precipitation and crystallisation steps.  相似文献   

17.
In situ ESR at 120–473 K permits to monitor formation of transient paramagnetic ions/complexes (isolated Pd+ sites; Pd+/H2O; Pd+/C6H6) upon interaction of isolated Pd2+ cations stabilized by the H-ZSM-5 matrix with different organic compounds and gas mixtures (NO, O2, H2O, H2, propene, benzene). The in situ study provides insight into the elementary steps of redox processes on isolated Pd species in H-ZSM-5 zeolite under realistic conditions. Adsorbed water stabilizes the transient paramagnetic complex and decreases the rate of Pd2+ to Pd0 reduction by H2. Strong bonding of NO x ligands to Pd2+ species suppresses the reduction of Pd(II) ions. Sorption of benzene on preoxidized Pd2+/HZSM-5 is accompanied by an easy formation of organic cation-radicals and of a Pd+/benzene complex, the paramagnetic Pd+/benzene structure indicating a surprisingly high resistance to further reduction to Pd0. Illumination of the Pd/HZSM-5 by UV-visible light causes no measurable change in the redox properties of the catalyst.  相似文献   

18.
A mechanism for the electrodeposition of acrylic resin on aluminium is proposed, based on experimental studies of acid value, anodic gas evaluation and anodic film resistance. The mechanism can be expressed as Alf Al3+ + 3e 2Al3+ + 3H2Of Al2O3 + 6H+ 2Al3+ + 6H2Of 2Al(OH)3 + 6H+ H+ + RCOC f RCOOH. This is different from the mechanism for zinc and steel, where it is metal ions from anodic dissolution which neutralize the macro-ions and cause a deposit on the anode surface.  相似文献   

19.
Gas phase reactions of Mo+ and W+ ions with the molecules of various oxidants (NO, O2, N2O, CH2O, C2H4O) were studied using ion cyclotron resonance. In oxidation with N2O the mono-, di- and trioxide metal cations are formed consecutively. The trioxide MO3 + ions of both metals react with CO to form CO2 and MO2 + ions. In this way, catalytic reaction N2O + CO N2 + CO2 occurs in the gas phase with MoO3 + /MoO2 + and WO3 +/WO2 + couples as catalysts. The rate constants have been measured for both stages of the catalytic cycle as well as for the stages of the catalyst preparation. Metal-oxygen bond energies were estimated for MoOx + and WOx + species with various x. The mechanism of CO oxidation with MoOx + and WOx + cations as catalysts in the gas phase is discussed in comparison with that for the oxidation over classical solid oxide catalysts.  相似文献   

20.
Polymerization of allyl methacrylate (AMA) with wool fabrics using different initiators, namely, potassium persulphate, Fe2+? H2O2, benzoyl peroxide, ceric ammonium nitrate, and vanadium pentanitrate, was investigated. The percent of polymer add-on depends upon the type and concentration of the initiator. Addition of metallic salts such as Fe3+ to the polymerization system enhances polymerization significantly when benzoyl peroxide and potassium persulphate are used independently as initiator. The opposite holds true for ceric ammonium nitrate and vanadium pentanitrate. With Fe2+? H2O2, on the other hand, the enhancement is marginal. Also studied was the incorporation of Li+, Cu++, and Fe3+ at different concentrations in AMA—wool–benzoyl peroxide polymerization systems. Determination of the polymer add-on on the basis of double bond analysis revealed that the remained double bond is governed by the magnitude of the polymer add-on as well as by the type of initiator.  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