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.
ZnO-based electrodes for one-step photocatalytic water splitting are designed by incorporating InN. The electronic and optical properties of (ZnO)1−x(InN)x alloys and ZnO with InN-like cluster formations ZnO:(InN)x are analyzed by means of first-principles approaches. We calculate the energy gaps Eg, the band-edge energies relative to the vacuum level, and the optical absorption, employing the GW0 method to describe single-particle excitations and the Bethe–Salpeter equation to model the two-particle exciton interactions. For ZnO and InN, the valence-band maximum (VBM) is EVBM ≈ −7.3 and −5.7 eV, and the energy gap is Eg ≈ 3.3 and 0.7 eV, respectively. Incorporating InN into ZnO, the random (ZnO)1−x(InN)x alloys up-shifts the VBM and down-shifts the conduction-band minimum (CBM). In addition, the presence of InN-like clusters enhances this effect and significantly narrows the band gap. For instance, the VBM and the energy gap for 12.5% InN are EVBM ≈ −6.5 and −6.1 eV, and Eg ≈ 2.2 and 1.9 eV for the alloy and the cluster structure, respectively. This impact on the electronic structure favors thus visible light absorption. With proper nanoclusters, the band edges straddle the redox potential levels of H+/H2 and O2/H2O, suggesting that ZnO–InN nanostructures can enhance the photocatalytic activity for overall solar-driven water splitting.  相似文献   

2.
The microstructures and phase composition of the pseudobinary ZrTi0.2V1.8 alloy were examined by scan electron microscope (SEM) and X-ray diffraction (XRD). Before hydrogenation, the hypoeutectic structure accompanied with ZrV2 + (ZrV2 + Zr) spherical-like texture has been observed in ZrTi0.2V1.8 and the dominant phase could be ascribed to the C15 Laves phase. Hydrogen absorption pressure–composition isotherms (PC isotherms) and hydriding kinetics of ZrTi0.2V1.8 were investigated by pressure reduction method using Sievert apparatus from 673 to 823 K. At hydrogen concentration 0.65 (H/A), the relative partial molar enthalpy and entropy calculated by Van’t Hoff equation are −60 ± 1 kJ mol−1 and −119 ± 1 J mol−1 K−1, respectively. In addition, two stages in the hydrogen absorption reaction between 673 and 823 K could be attributed to the different hydrogen absorption mechanisms including redistribution of the hydrogen atoms in the hydride phase and the diffusion of hydrogen in the β-phase. The activation energy Ea of the alloy is ∼3.6 kJ mol−1 for the first absorption stage and ∼61.9 kJ mol−1 for the second one.  相似文献   

3.
A series of equivalent substitution solid-solution Na(BixTa1−x)O3 (x = 0–0.10) was prepared by a simple hydrothermal method using Ta2O5 and NaBiO3 as precursors. The Na(Bi0.08Ta0.92)O3 photocatalyst exhibited the highest performance of H2 evolution (59.48 μmol h−1 g−1) under visible-light irradiation (λ > 400 nm) without co-catalyst, whereas no H2 evolution is observed for NaTaO3 under the same conditions. The UV-Vis spectra indicate that the Na(Bi0.08Ta0.92)O3 powders can absorb not only ultraviolet light like pure NaTaO3 powder but also the visible-light spectrum. The absorption edge corresponds to a band gap of 2.88 eV. The results of density functional theory calculation illuminate that the visible-light absorption bands in the Na(BixTa1−x)O3 catalysts are attributed to the band transition from the O2p to the Bi2s + 2p + Ta5d hybrid orbital.  相似文献   

4.
Through the use of a water balance experiment, the electro-osmotic drag coefficients of Nafion 115 were obtained under several conditions (as a function of water content and thermodynamics conditions). For the cases when the anode was fully hydrated (corresponding to water content λ ≈ 14 in the adjacent membrane) and the cathode suffered from drying when dry air was supplied (λ ≈ 2), the electro-osmotic drag coefficients varied from 0.82 (±0.06) to 0.50 (±0.03) H2O/H+ when the current density varied from 0.4 to 1.0 A cm−2 (95% confidence level). When the current density increased, the electro-osmotic drag coefficient decreased. When the water content at the anode increased from λ ≈ 5 to λ ≈ 14, the cathode was supplied with dry air (λ ≈ 2), and the fuel cell discharged constant current density at 0.6 A cm2, the electro-osmotic drag coefficient increased from 0.44 (±0.06) to 0.68 (±0.06) H2O/H+ (95% confidence level). Higher relative humidity gas leads to a higher electro-osmotic drag coefficient at constant current density.  相似文献   

5.
Doping, compensation and photovoltaic performance have been investigated in all-metal-organic vapour-phase deposition (MOCVD) grown CdTe/CdS solar cells that were co-doped with arsenic and chlorine.Although arsenic chemical concentration is in the range of 1017-1.5×1019 cm−3, the maximum net acceptor concentration is only in the order of 1014 cm−3, as determined by capacitance-voltage characteristics. Admittance spectroscopy revealed shallow traps at 0.055 eV which were attributed to AsTe; its compensation by Cdi is discussed. Formation of the alloy CdSxTe1−x is linked to deep levels at EV+∼0.55 eV and EV+∼0.65 eV. Limits to the diffusion of photo-generated carriers were considered to be important in determining photovoltaic performance rather than carrier lifetime. Prospects for optimizing the performance of such co-doped MOCVD-grown devices are discussed.  相似文献   

6.
We have made an attempt to evaluate the variation in the electron discharge (ED) pattern of anaerobic consortia as a function of pretreatment viz., chemical, heat-shock, acid and oxygen-shock in comparison with untreated mixed consortia during fermentative hydrogen (H2) production. Experiments were performed with dairy wastewater as substrate using anaerobic mixed consortia as biocatalyst (pretreated individually and in combination). Cyclic voltammetry (CV) elucidated significant variation in the ED pattern of mixed consortia along with H2 production and substrate degradation (SD) as a function of pretreatment method applied. Higher ED was observed with all pretreated consortia which can be attributed to the stable proton (H+) shuttling due to the suppression of methanogenic activity. Oxygen-shock method and untreated consortia showed lower H2 production and higher SD among the variations studied, while, combined pretreated consortia resulted higher H2 production and lower SD. Lower ED observed with untreated consortia suggests the H+ reduction during methanogenesis rather than the inter-conversion of metabolites, which is presumed to be necessary for H2 production. ED observed with combined pretreated consortia corroborated well with the observed H2 production. Redox pairs were visualized on the voltammograms with almost all the experimental variations studied except untreated consortia. The potentials (E0) of redox pairs observed were corresponding to intracellular electron carriers viz., NAD+/NADH (E0 −0.32 V) and FAD+/FADH2 (E0 −0.24 V).  相似文献   

7.
New low-band-gap copolymers, including thieno[3,2-b]thiophene and 2,1,3-benzothiadiazole, were synthesized as photovoltaic materials. Thiophene was introduced to provide extended π-conjugation length and charge transfer properties. A band gap (Egop=1.62 eV, Egec=1.51 eV) of this polymer was investigated through UV-vis spectroscopy and cyclic voltammetry. A bulk heterojunction structure of glass/indium tin oxide (ITO)/PEDOT:PSS/polymer-PCBM(1:3)/LiF/Al was fabricated for investigating photovoltaic properties. PC71BM was used as an acceptor material, due to its increased absorption in the visible region, in comparison with PC61BM. In this polymer, incident photon-to-current conversion efficiency (IPCE) was as high as 50%. Moreover, maximum power conversion efficiency (PCE) of up to 1.72% was achieved under AM 1.5 G conditions. It demonstrated relatively high VOC (0.67 V) and JSC (6.86 mA/cm2), while a low fill factor value (0.37) was obtained.  相似文献   

8.
We prepared Ti1.4V0.6Ni ribbons by arc-melting and subsequent melt-spinning techniques. Ti1.4V0.6Ni + x Mg (x = 1, 1.5, 2, 2.5 and 3, wt.%) composite alloys were obtained by the mechanical ball-milling method. The structures and hydrogen storage properties of alloys were investigated. Ti1.4V0.6Ni + x Mg composite alloys contained icosahedral quasicrystalline phase, Ti2Ni-type phase, β-Ti solid-solution phase and metallic Mg. The electrochemical and gaseous hydrogen storage properties of alloys were improved with Mg addition. Ti1.4V0.6Ni + 2 Mg alloy showed maximum electrochemical discharge capacity of 282.5 mAh g−1 as well as copacetic high-rate discharge ability of 82.3% at the discharge current density of 240 mA g−1 compared with that of 30 mA g−1, and the cycling life achieved above 200 mAh g−1 after 50 consecutive cycles of charging and discharging. The hydrogen absorption/desorption properties of Ti1.4V0.6Ni + x Mg (x = 1, 2 and 3, wt.%) alloys were better than Ti1.4V0.6Ni. Ti1.4V0.6Ni + 3 Mg alloy also exhibited a favorable hydrogen absorption capacity of 1.53 wt.%. The improvement in the hydrogen storage characteristics caused by adding Mg may be ascribed to better hydrogen diffusion and anti-corrosion ability.  相似文献   

9.
Here, novel core/shell polydopamine@Ni-MOF (pDA@Ni-MOF) heterogeneous nanostructures are synthesized via a simple one-pot nucleation-growth technique. This rational core/shell design method provide a uniform Ni-MOF shell thickness (shell: ~ 10 nm) as well as homogeneous wrapping of pDA templates with quite narrow size distributions. The obtained band properties of bare pDA (ECB = ?0.35 eV and EVB = 2.95 eV vs normal hydrogen electrode (NHE)) and bare Ni-MOF (ECB = ?0.49 eV and EVB = 2.85 eV vs NHE) clearly revealed charge separation is occurred on pDA by absorbing light due to π-π1 transition, and photogenerated electrons on conduction band (CB) of pDA was migrated to CB of Ni-MOF. Specifically, the photoelectrochemical (PEC) water performance of pDA@Ni-MOF photoanodes with highest current density is recorded as 8.61 mA/cm2 at 0.77 V vs. RHE under visible LED irradiation, which is significantly higher than bare pDA (0.008 V vs. RHE) and bare Ni-MOF (0.011 V vs. RHE) at the same conditions. Note that, the higher photon absorption properties of pDA in core together with high interaction valence bond between two semiconductors could generate electron rich state giving rise to faster electron transfer kinetics as next generation of MOF based hybrid materials with regular morphologies.  相似文献   

10.
In an attempt to identify an active material for use in lithium secondary batteries with high energy density, we investigated the electrochemical properties of gallium (III) sulfide (Ga2S3) at 30 °C. Ga2S3 shows two sloping plateaus in the potential range between 0.01 V and 2.0 V vs. (Li/Li+). The specific capacity of the Ga2S3 electrode in the first delithiation is ca. 920 mAh g−1, which corresponds to 81% of the theoretical capacity (assuming a 10-electron reaction). The capacity in the 10th cycle is 63% of the initial capacity. Ex situ X-ray diffraction and X-ray absorption fine structure analyses revealed that the reaction of the Ga2S3 electrode proceeds in two steps: Ga2S3 + 6Li+ + 6e ? 2Ga + 3Li2S and Ga + xLi+ + xe ? LixGa.  相似文献   

11.
Spatial, quantitative, and temporal information regarding the water content distribution in the transverse-plane between the catalyst layers of an operating polymer-electrolyte membrane fuel cell (PEMFC) is essential to develop a fundamental understanding of water dynamics in these systems. We report 1H micro-magnetic resonance imaging (MRI) experiments that measure the number of water molecules per SO3H group, λ, within a Nafion®-117 membrane between the catalyst stamps of a membrane-electrode assembly, MEA. The measurements were made both ex situ, and inside a PEMFC operating on hydrogen and oxygen. The observed 1H MRI T2 relaxation time of water in the PEM was measured for several known values of λ. The signal intensity of the images was then corrected for T2 weighting to yield proton density-weighted images, thereby establishing a calibration curve that correlates the 1H MRI density-weighted signal with λ. Subsequently, the calibration curve was used with proton density weighted (i.e., T2-corrected) signal intensities of transverse-plane 1H MRI images of water in the PEM between the catalyst stamps of an operating PEMFC to determine λ under various operational conditions. For example, the steady state, transverse-plane λ was 9 ± 1 for a PEMFC operating at ∼26.4 mW cm−2 (∼20.0 mA, ∼0.661 V, 20 °C, flow rates of the dry H2(g) and O2(g) were 5.0 and 2.5 mL min−1, respectively).  相似文献   

12.
To improve upon our previously reported slow hydrogen evolution rate RH at the energy-efficient lower voltages in CAWE (carbon-assisted water electrolysis) at room temperature, new results using different carbons and catalysts to improve RH are reported here. Compared to earlier results with carbon GX203, about a ten-fold increase in RH is reported using high surface area carbon BP2000 at the operating voltage Eo = 1.12 V. With added FeSO4 catalyst, Eo is lowered to 0.72 V without lowering RH, representing about 30% decrease in the energy barrier of the process. For comparison, in water electrolysis without carbon, measurable RH is observed only for Eo ≥ 2 V. This large improvement in RH at the energy efficient Eo = 0.72 V is suggested to result from nanoscale particle size of carbon BP2000 as well as from electrons provided by the catalyst through the reaction Fe2+ ? Fe3+ + e. By measuring the amounts of H2 evolved at the cathode and CO2 evolved at the anode using gas chromatography, the mechanism for CAWE is established to be the reaction: C (s) + 2H2O (?) → CO2 (g) + 2H2 (g). The reaction slows down with time as carbon is depleted by oxidation.  相似文献   

13.
Single-electron transfer (SET) between donor-acceptor complex PTZ–TCBQ (phenothiazine:tetrachloro-p-benzoquinone = 1/1) and thiol-stabilized gold nanoclusters Au25(SC2H4Ph)18TOA+ (abbreviated as Au25) is firstly researched here. Cyclic voltammetry (CV), UV–vis spectroscopy, electron spin resonance (ESR) and nuclear magnetic resonance (NMR) are utilized to study their electron transfer reaction. The characteristic results show only one-electron transfer from Au25 to the complex, and the excess amount of PTZ–TCBQ complex cannot oxidate Au250 further as 1HNMR observed. So those results indicate that this well defined gold nanoclusters with a molecular-like redox behavior as an electron donor can be applied as a solar photocatalytic nanomaterial, SET catalyst and so on.  相似文献   

14.
We examine further the electrochemical oxidation of carbon in molten carbonate, based on analysis of published research. Ascending and descending branches of voltage hysteresis found in current sweeps of atomically-ordered graphite and of disordered carbon (coal char) are separated by about 0.20–0.25 V and by 0.10–0.15 V for ordered and disordered forms, respectively, over a wide band of current density, 0.03–0.10 A/cm2. The higher voltage of the descending branch is in rough agreement with prediction of the Y. Li model for the carbon/carbonate electrode in the same current range, for ordered graphite (La = 70–100 nm) and for disordered structures (La = 3–5 nm), respectively. We suggest that the amplitude of the hysteresis represents the difference between the overvoltage requirements for 2- and 4 electron net transfer processes, respectively. The 2 e− reaction (C + CO32− = CO + CO2 + 2e) dominates the low current segment (LCS) of our previous analysis, and the more hindered 4e− transfer reaction (C + 2CO32− = 3CO2 + 4e) dominates the high current segment (HCS). The voltage increase separating LCS from HCS is effected by accumulation of CO2 within small, melt-filled pores to form highly supersaturated solutions of CO2, which enhance anode voltage by a concentration overpotential of 0.10–0.25 V. Overpotential increases with reaction extent until (1) overall polarization inhibits the interior reaction and shifts CO2 production to the more accessible exterior surface, or, (2) at a critical concentration (dependent on surface tension and pore diameter) bubbles nucleate and block current flow in the pores. Further support for this picture comes from the often-reported deviation of the gas composition from the CO/CO2 ratio of the Boudouard equilibrium at atmospheric pressure, as open circuit conditions are approached in an electrochemical cell. Our interpretation accounts for the mole fraction of CO2 at open circuit being greater than predicted from the Boudouard equilibrium.  相似文献   

15.
The storage of H2 molecules was studied theoretically on charged and uncharged Mg decorated graphene surfaces using density-functional theory and by incorporating the van der Waals (vdW) interactions. We found that an increase in the number of Mg atoms and H2 molecule increases the net interaction of the hydrogen molecule with the surface. The Mg-Gr+ has the hydrogen storage capacity of up to nine H2 molecules, with the average adsorption energy of −0.134 eV/H2. Also, we found that hydrogen molecules play an important role in the interaction between the graphene surface and the Mg atom. The charge density difference analysis showed that electron transfer occurs from H2 molecules to Mg atom in uncharged system. However, the Bader charge analysis showed that the positive charges in the Mg-Gr+ and nH2-Mg-Gr+ systems are concentrated on the Mg atom. When the number of H2 molecules reaches more than 4, the charge transfer instead occurs from the Mg atom to H2 molecules as well as to the graphene surface. This results in better interaction between the Mg atom and the Gr+ surface.  相似文献   

16.
xLiH + M composites, where M = Mg or Ti, are suggested as new candidates for negative electrode for Li-ion batteries. For this purpose, the xLiH + M electrode is prepared using the mechanochemical reaction: MHx + xLi → xLiH + M or by simply grinding a xLiH + M mixture. The most promising electrochemical behaviour is obtained with the (2LiH + Mg) composite prepared via a mechanochemical reaction between MgH2 and metallic Li leading to a very divided composite in which Mg crystallites of 20 nm size are embedded in a LiH matrix. Reversible capacities of 1064 mAh g−1 (three times as much as the one of graphite) and 600 mAh g−1 are reached for these phase mixtures after 1 and 28 h of grinding in vertical and planetary mill, respectively. The (2LiH + Ti) mixture prepared via the mechanochemical reaction between TiH2 and Li exhibits a reversible capacity of 428 mAh g−1. From X-ray diffraction measurements, the performances of the electrodes are attributed to the electrochemical conversion reaction: M + xLiH ↔ MHx + xLi+ + xe (M = Mg, Ti) followed for M = Mg by an alloying process where M reacts with lithium ions to form Mg1−xLix alloys.  相似文献   

17.
Hydrogen (1H) trapped at intermetallic particles (IPs) in an aluminum alloy, 6061-T6, was visualized with secondary ion mass spectrometry (SIMS) by precisely excluding the false signal which is caused by background hydrogen (HBG). The interference of the HBG was avoided by a unique continuous pre-sputtering (pre-digging) by a primary ion beam of SIMS into a sample in combination with silicon sputtering prior to the SIMS measurement of the sample and we succeeded in visualizing the exact signal of 1H trapped by IPs at subsurface layer of the sample charged in high-pressure hydrogen gas. The thermal desorption analysis clarified that the desorption energy (Ed) of the IPs was 200 kJ/mol or higher, which was extremely higher than Ed for lattice interstice, dislocations, and vacancies. High density hydrogen was concentratedly trapped at IPs in the subsurface layer in contact with the hydrogen gas. This nature causes an extremely low effective hydrogen diffusivity of 6061-T6 of the order of 10?14 m2/s even at 200 °C and may eventually give a high HE resistance to 6061-T6.  相似文献   

18.
The ternary [Li+]0.09[MePrPyr+]0.41[NTf2]0.50 room temperature ionic liquid was obtained by dissolution of solid lithium bis(trifluoromethanesulfonyl)imide (LiNTf2) in liquid N-methyl-N-propylpyrrolidinium bis(trifluoromethanesulfonyl)imide ([MePrPyr+][NTf2]), and studied as an electrolyte for lithium-ion batteries. The graphite-lithium (C6Li) anode, working together with vinylene carbonate as an additive showed ca. 90% of its initial discharge capacity after 50 cycles. The addition of vinylene carbonate to the neat ionic liquid results in the formation of the protective coating (SEI) on both the lithium and graphite anodes. The SEI formation increases the rate of the charge transfer reaction as well as protects the anode from chemical passivation (corrosion). The graphite-lithium (C6Li) anode shows good cyclability and Coulombic efficiency in the presence of 10 wt.% of vinylene carbonate as an additive to the ionic liquid.  相似文献   

19.
In this study, new electrolytes for Li-ion batteries in the form of lithium salt solutions in room temperature imidazolium ionic liquids (RTIL) are reported. The ionic liquids applied, for higher reduction potential stability, were substituted at position C2 with oligooxyethylene groups of various length ([Im nEO]+X; where: n = 0, 3, 7, 20 and X = Cl, BF4, N(CF3SO2)2). It was found that they are good solvents for lithium salts (LiBF4, LiN(CF3SO2)2, {[CH3(OCH2CH2)3O]3BC4H9}Li) forming liquid solutions of low glass transition temperature (Tg in the −70 to −40 °C range). Ionic conductivity depends on the length of oxyethylene substituent in Im nEO and on the concentration of the salt applied, for 10 mol%, σRT is of the order of 10−4 S cm−1. On the basis of polarization measurements by the variable-current method, the proportion of lithium cations in electric charge transfer (t+) has been determined. The values obtained (typical for ionic liquids) are low and depend on n and lithium salt concentration but do not exceed a dozen or so percent.  相似文献   

20.
This paper reports an experimental study of the degradation of single PBI-based high temperature MEAs doped with phosphoric acid. The study is carried out by operating the single MEAs for long periods in steady state, the degradation is quantified considering the voltage decay rate. Besides the most common operating condition suggested by the MEAs producer (T = 160 °C, i = 0.2 A cm−2, λH2=1.2λH2=1.2, λair = 2), the study also investigates higher operating temperature (T = 180 °C), higher current density (i = 0.4 A cm−2) and double air flow rate (λair = 4). A temperature of 180 °C accelerates the degradation of the MEA which increases from around 8 μV h−1 up to around 19 μV h−1. On the opposite side, operating the MEA at i = 0.4 A cm−2 reduces the voltage degradation rate down to 4 μV h−1 and increases the power output making this condition particularly interesting. EIS, CV and LSV are used to clarify the causes of degradation. A consistent increase in the charge transfer resistance is observed and is related to the loss of catalyst active area due to catalyst agglomeration, carbon corrosion and possible acid leaching. Concerning the electrolyte membrane, a slight decrease in the proton conductivity is measured, a major effect on degradation is played by the increasing gas crossover rate and by the short circuit current.  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