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.
Nanocrystalline Ce0.8Sm0.2O1.9 (SDC) has been synthesized by a combined EDTA–citrate complexing sol–gel process for low temperature solid oxide fuel cells (SOFCs) based on composite electrolyte. A range of techniques including X-ray diffraction (XRD), and electron microscopy (SEM and TEM) have been employed to characterize the SDC and the composite electrolyte. The influence of pH values and citric acid-to-metal ions ratios (C/M) on lattice constant, crystallite size and conductivity has been investigated. Composite electrolyte consisting of SDC derived from different synthesis conditions and binary carbonates (Li2CO3–Na2CO3) has been prepared and conduction mechanism is discussed. Water was observed on both anode and cathode side during the fuel cell operation, indicating the composite electrolyte is co-ionic conductor possessing H+ and O2− conduction. The variation of composite electrolyte conductivity and fuel cell power output with different synthesis conditions was in accordance with that of the SDC originated from different precursors, demonstrating O2− conduction is predominant in the conduction process. A maximum power density of 817 mW cm−2 at 600 °C and 605 mW cm−2 at 500 °C was achieved for fuel cell based on composite electrolyte.  相似文献   

2.
Pure hydrogen can be stored and supplied directly to polymer electrolyte fuel cell by the redox of iron oxide: Fe3O4 + 4H2 → 3Fe + 4H2O and 4H2O + 3Fe → Fe3O4 + 4H2. Four bimetal-modified samples were prepared by impregnation. The hydrogen storage properties of the samples were investigated. The result shows that the Fe2O3–Mo–Al sample presented the most excellent catalytic activity and cyclic stability. H2 forming temperature and H2 forming rate could be surprisingly decreased and enhanced, respectively. The average H2 forming temperature at the rate of 250 μmol min−1·Fe-g−1 for Fe2O3–Mo–Al in the first 4 cycles could be decreased from 469 °C before the addition of Mo–Al to 273 °C after the addition of Mo–Al. The reason for it may be that the Mo–Al additive in the sample can prevent from the sintering of the particles and accelerate the H2O decomposition due to Mo taking part in the redox reaction. The average storage capacity of Fe2O3–Mo–Al was up to 4.68 wt%.  相似文献   

3.
The novel core–shell nanostructured SDC/Na2CO3 composite has been demonstrated as a promising electrolyte material for low-temperature SOFCs. However, as a nanostructured material, stability might be doubted under elevated temperature due to their high surface energy. So in order to study the thermal stability of SDC/Na2CO3 nanocomposite, XRD, BET, SEM and TGA characterizations were carried on after annealing samples at various temperatures. Crystallite sizes, BET surface areas, and SEM results indicated that the SDC/Na2CO3 nanocomposite possesses better thermal stability on nanostructure than pure SDC till 700 °C. TGA analysis verified that Na2CO3 phase exists steadily in the SDC/Na2CO3 composite. The performance and durability of SOFCs based on SDC/Na2CO3 electrolyte were also investigated. The cell delivered a maximum power density of 0.78 W cm−2 at 550 °C and a steady output of about 0.62 W cm−2 over 12 h operation. The high performances together with notable thermal stability make the SDC/Na2CO3 nanocomposite as a potential electrolyte material for long-term SOFCs that operate at 500–600 °C.  相似文献   

4.
The composite LaNi3.7Al1.3/Ni–S–Co alloy film was prepared by molten salt electrolysis and aquatic electrodeposition orderly. With Na3AlF6–La2O3–Al2O3 (91:8:1) system as molten salt electrolyte, the LaNi3.7Al1.3 alloy film was obtained by galvanostatic electrolysis at 100 mA cm−2. The results showed that the La3+ and Al3+ ions could be co-reduced on the nickel cathode to form LaNi3.7Al1.3 film, i.e. La3+ + 1.3Al3+ + 6.9e + 3.7Ni = LaNi3.7Al1.3 at c.a. −0.5 V, which is much lower than that of the theoretical decomposition potential of lanthanum and aluminum. With high HER activity, the composite LaNi3.7Al1.3/Ni–S–Co film (η150 = 65 mV, 353 K) could absorb large amount of H atoms, which would be oxidized and therefore effectively avoid the dissolution of the Ni–S–Co film under the state of open-circuit and consequently prolong the lifetime of the cathode.  相似文献   

5.
On the basis of extreme similarity between the triangle phase diagrams of LiNiO2–LiTiO2–Li[Li1/3Ti2/3]O2 and LiNiO2–LiMnO2–Li[Li1/3Mn2/3]O2, new Li–Ni–Ti–O series with a nominal composition of Li1+z/3Ni1/2−z/2Ti1/2+z/6O2 (0 ≤ z ≤ 0.5) was designed and attempted to prepare via a spray-drying method. XRD identified that new Li–Ni–Ti–O compounds had cubic rocksalt structure, in which Li, Ni and Ti were evenly distributed on the octahedral sites in cubic closely packed lattice of oxygen ions. They can be considered as the solid solution between cubic LiNi1/2Ti1/2O2 and Li[Li1/3Ti2/3]O2 (high temperature form). Charge–discharge tests showed that Li–Ni–Ti–O compounds with appropriate compositions could display a considerable capacity (more than 80 mAh g−1 for 0.2 ≤ z ≤ 0.27) at room temperature in the voltage range of 4.5–2.5 V and good electrochemical properties within respect to capacity (more than 150 mAh g−1 for 0 ≤ z ≤ 0.27), cycleability and rate capability at an elevated temperature of 50 °C. These suggest that the disordered cubic structure in some cases may function as a good host structure for intercalation/deintercalation of Li+. A preliminary electrochemical comparison between Li1+z/3Ni1/2−z/2Ti1/2+z/6O2 (0 ≤ z ≤ 0.5) and Li6/5Ni2/5Ti2/5O2 indicated that charge–discharge mechanism based on Ni redox at the voltage of >3.0 V behaved somewhat differently, that is, Ni could be reduced to +2 in Li1+z/3Ni1/2−z/2Ti1/2+z/6O2 while +3 in Li6/5Ni2/5Ti2/5O2. Reduction of Ti4+ at a plateau of around 2.3 V could be clearly detected in Li1+z/3Ni1/2−z/2Ti1/2+z/6O2 with 0.27 ≤ z ≤ 0.5 at 50 °C after a deep charge associated with charge compensation from oxygen ion during initial cycle.  相似文献   

6.
Several substituted titanates of formula Li4−xMgxTi5−xVxO12 (0 ≤ x ≤ 1) were synthesized (and investigated) as anode materials in rechargeable lithium batteries. Five samples labeled as S1–S5 were calcined (fired) at 900 °C for 10 h in air, and slowly cooled to room temperature in a tube furnace. The structural properties of the synthesized products have been investigated by X-ray diffraction (XRD), scanning electron microscope (SEM) and Fourier transmission infrared (FTIR). XRD explained that the crystal structures of all samples were monoclinic while S1 and S3 were hexagonal. The morphology of the crystal of S1 was spherical while the other samples were prismatic in shape. SEM investigations explained that S4 had larger grain size diameter of 15–16 μm in comparison with the other samples. S4 sample had the highest conductivity 2.452 × 10−4 S cm−1. At a voltage plateau located at about 1.55 V (vs. Li +), S4 cell exhibited an initial specific discharge capacity of 198 mAh g−1. The results of cyclic voltammetry for Li4−xMgxTi5−xVxO12 showed that the electrochemical reaction was based on Ti4+/Ti3+ redox couple at potential range from 1.5 to 1.7 V. There is a pair of reversible redox peaks corresponding to the process of Li+ intercalation and de-intercalation in the Li–Ti–O oxides.  相似文献   

7.
Phase evolution, structure, thermal property, morphology, electrical property and reactivity of a perovskite-type cathode system, La0.75Sr0.25 Mn0.95−xCoxNi0.05O3+δ (0.1 ≤ x ≤ 0.3), are reported. The samples are synthesized using metal acetates by the Pechini method. A perovskite-type phase is formed after calcination at ∼700 °C and a rhombohedral symmetry of R – 3c space group is stabilized at ∼1100 °C. An increase in x decreases the unit cell volume linearly, accompanying with a linear decrease in bond lengths and tilt angle. The differential thermal analysis suggests the phase stabilization for a temperature range, 50–1100 °C. The thermo-gravimetric, thermal expansion, and electrical and ionic conductivities studies suggest presence of a Jahn–Teller transition at ∼260–290 °C. The samples with x = 0.1 mol exhibit electrical conductivity of ∼55 S cm−1 at ∼600 °C, activation energy of ∼0.13 eV, coefficient of thermal expansion of ∼12 × 106 °C−1, crystallite size of ∼45 nm, Brunauer–Emmett–Teller (BET) surface area of 1.26 m2 g−1 and average particle size of ∼0.9 μm. A fairly high ionic conductivity, 5–9 × 10−2 S cm−1 makes the sample with x = 0.1 mole suitable for intermediate-temperature solid oxide fuel cell cathode applications. The experimental results are discussed with the help of the defect models proposed for La1−xSrxMnO3+δ.  相似文献   

8.
Porous nanostructured LiFePO4 powder with a narrow particle size distribution (100–300 nm) for high rate lithium-ion battery cathode application was obtained using an ethanol based sol–gel route employing lauric acid as a surfactant. The synthesized LiFePO4 powders comprised of agglomerates of crystallites <65 nm in diameter exhibiting a specific surface area ranging from 8 m2 g−1 to 36 m2 g−1 depending on the absence or presence of the surfactant. The LiFePO4 obtained using lauric acid resulted in a specific capacity of 123 mAh g−1 and 157 mAh g−1 at discharge rates of 10C and 1C with less than 0.08% fade per cycle, respectively. Structural and microstructural characterization were performed using X-ray diffraction (XRD), scanning electron microscopy (SEM) and high-resolution transmission electron microscopy (HRTEM) with energy dispersive X-ray (EDX) analysis while electronic conductivity and specific surface area were determined using four-point probe and N2 adsorption techniques.  相似文献   

9.
The CO inhibition effect on H2 permeance through commercial Pd-based membranes was analysed by means of permeation measurements at different CO compositions (0–30% molar) and temperatures (593–723 K) with the aim to determine the increase of the membrane area in order to compensate the H2 flux reduction owing to the CO inhibition effect. The permeance of H2 fed with carbon monoxide was observed to decrease with respect to the case of pure hydrogen. At 647 K the H2 permeance of a pure feed of 316 μmol m−2 s−1 Pa−0.5 reduces progressively until 275 μmol m−2 s−1 Pa−0.5 when 15% or more of CO is present in the system, until it reaches a plateau at 20%. The inhibition effect occurring when CO is present in the feed stream reduces with the progressive temperature increase; the reduction of the permeance decreases exponentially by 23% at 593 K and by 3% at 723 K with 10% of CO. The inhibition effect is seen to be reversible. An H2 flux profile in a Sieverts' plot shows the effect produced by the increase of the CO composition along the Pd-based membrane length. The H2 flux profile allows the area of a Pd-based membrane to be evaluated in order to have the same permeate flow rate of H2 when it is fed with CO or as a pure stream. Moreover, a qualitative comparison between the H2 flux profiles and a previously proposed model has been carried out.  相似文献   

10.
A new anode material for intermediate temperature solid oxide fuel cells (IT-SOFCs) with a composite of La0.7Sr0.3Cr1−xNixO3 (LSCN), CeO2 and Ni has been synthesized. EDX analysis showed that 1.19 at% Ni was doped into the perovskite-type La0.7Sr0.3CrO3 and Ce could not be detected in the perovskite phases. Results showed that the fine CeO2 and Ni were highly dispersed on the La0.7Sr0.3Cr1−xNixO3 substrates after calcining at 1450 °C and reducing at 900 °C. The thermal expansion coefficient (TEC) of the as-prepared anode material is 11.8 × 10−6 K−1 in the range of 30–800 °C. At 800 °C, the electrical conductivity of the as-prepared anode material calcined at 1450 °C for 5 h is 1.84 S cm−1 in air and 5.03 S cm−1 in an H2 + 3% H2O atmosphere. A single cell with yttria-stabilized zirconia (YSZ, 8 mol% Y2O3) electrolyte and the new materials as anodes and La0.8Sr0.2MnO3 (LSM)/YSZ as cathodes was assembled and tested. At 800 °C, the peak power densities of the single cell was 135 mW cm−2 in an H2 + 3% H2O atmosphere.  相似文献   

11.
The gas permeability of H2S gas at 150 °C through ultra-thin cesium hydrogen sulfate (CsHSO4) membranes has been investigated. Gas chromatography–mass spectrometry analyses indicate that CsHSO4 membrane is impermeable to H2S gas under test conditions. The apparent micropore diameter of the membrane averaged between 9.5 and 11.5 Å with a maximum permeance of 0.09 Barrer (6.75 × 10−19 m2 s−1 Pa−1). Atomic force microscope and X-ray diffraction analyses show respectively that the surface morphology and crystal structure of the membranes are preserved, with no adverse effect from prolonged exposure to H2S gas. Electrochemical impedance spectroscopy analysis confirm over a 30% decrease in membrane resistance via an 80% reduction in membrane thickness.  相似文献   

12.
The all-solid-state Li–In/Li4Ti5O12 cell using the 80Li2S·20P2S5 (mol%) solid electrolyte was assembled to investigate rate performances. It was difficult to obtain the stable performance at the charge current density of 3.8 mA cm−2 in the all-solid-state cell. In order to improve the rate performance, the pulverized Li4Ti5O12 particles were applied to the all-solid-state cell, which retained the reversible capacity of about 90 mAh g−1 at 3.8 mA cm−2. The 70Li2S·27P2S5·3P2O5 glass–ceramic, which exhibits the higher lithium ion conductivity than the 80Li2S·20P2S5 solid electrolyte, was also used. The Li–In/70Li2S·27P2S5·3P2O5 glass–ceramic/pulverized Li4Ti5O12 cell was charged at a current density higher than 3.8 mA cm−2 and showed the reversible capacity of about 30 mAh g−1 even at 10 mA cm−2 at room temperature.  相似文献   

13.
Flame spray synthesis (FSS), a large-scale powder processing technique is used to prepare nanoscale La0.6Sr0.4CoO3−δ powder for solid oxide fuel cell cathodes from water-based nitrate solutions. Influence of processing is investigated on basis of the as-synthesised powders by X-ray powder diffraction (XRD), thermal gravimetric analysis (TGA), nitrogen adsorption (BET) and electron microscopy (SEM and TEM).Against the background of a nanostructured cathode morphology for an intermediate temperature solid oxide fuel cell (IT-SOFC) at 600 °C, an optimised and high surface area flame-made La0.6Sr0.4CoO3−δ nanopowder of 29 m2 g−1 is used to investigate its performance and chemical reaction with common electrolytes (Y0.16Zr0.84O2−δ, Ce0.9Gd0.1O2−δ and Sc0.20Ce0.01Zr0.79O2−δ). Secondary phase analysis from XRD measurements revealed a substantially lower La2Zr2O7 and SrZrO3 formation in comparison to conventional spray pyrolysed and submicron powder of about 9 m2 g−1. TGA and resistivity measurements proofed that La0.6Sr0.4CoO3−δ is non-sensitive towards carbonate formation under CO2 containing atmospheres. Electronic bulk conductivity of 2680 S cm−1 (600 °C) and 3340 S cm−1 (500 °C) were measured in air and as function of oxygen partial pressure (2 × 105 Pa > p(O2) > 1.2 × 10−2 Pa) in the temperature range between 400 and 900 °C. Electrochemical performance is determined by impedance spectroscopy on symmetrical cells of screen printed nanoscale La0.6Sr0.4CoO3−δ on Ce0.9Gd0.1O2−δ substrates from which an area specific resistance (ASR) of 0.96 Ω cm2 at 600 °C and 0.14 Ω cm2 at 700 °C were obtained.  相似文献   

14.
Microalgae present some advantageous qualities for reducing carbon dioxide (CO2) emissions from ethanol biorefineries. As photosynthetic organisms, microalgae utilize sunlight and CO2 to generate biomass. By integrating large-scale microalgal cultivation with ethanol biorefineries, CO2 sequestration can be coupled with the growth of algae, which can then be used as feedstock for biodiesel production. In this case study, a 50-mgy ethanol biorefinery in Iowa was evaluated as a candidate for this process. Theoretical projections for the amount of land needed to grow algae in raceway ponds and the oil yields of this operation were based on the amount of CO2 from the ethanol plant. A practical algal productivity of 20 g m−2 d−1 would require over 2,000 acres of ponds for complete CO2 abatement, but with an aggressive productivity of 40–60 g m−2 d−1, a significant portion of the CO2 could be consumed using less than 1,000 acres. Due to the cold temperatures in Iowa, a greenhouse covering and a method to recover waste heat from the biorefinery were devised. While an algal strain, such as Chlorella vulgaris, would be able to withstand some temperature fluctuations, it was concluded that this process is limited by the amount of available heat, which could maintain only 41 acres at 73 °F. Additional heating requirements result in a cost of 10–40 USD per gallon of algal oil, which is prohibitively expensive for biodiesel production, but could be profitable with the incorporation of high-value algal coproducts.  相似文献   

15.
The initialization of an anode-supported single-chamber solid-oxide fuel cell, with NiO + Sm0.2Ce0.8O1.9 anode and Ba0.5Sr0.5Co0.8Fe0.2O3−δ + Sm0.2Ce0.8O1.9 cathode, was investigated. The initialization process had significant impact on the observed performance of the fuel cell. The in situ reduction of the anode by a methane–air mixture failed. Although pure methane did reduce the nickel oxide, it also resulted in severe carbon coking over the anode and serious distortion of the fuel cell. In situ initialization by hydrogen led to simultaneous reduction of both the anode and cathode; however, the cell still delivered a maximum power density of ∼350 mW cm−2, attributed to the re-formation of the BSCF phase under the methane–air atmosphere at high temperatures. The ex situ reduction method appeared to be the most promising. The activated fuel cell showed a peak power density of ∼570 mW cm−2 at a furnace temperature of 600 °C, with the main polarization resistance contributed from the electrolyte.  相似文献   

16.
Nano-CuCo2O4 is synthesized by the low-temperature (400 °C) and cost-effective urea combustion method. X-ray diffraction (XRD), high resolution transmission electron microscopy (HR-TEM) and selected area electron diffraction (SAED) studies establish that the compound possesses a spinel structure and nano-particle morphology (particle size (10–20 nm)). A slight amount of CuO is found as an impurity. Galvanostatic cycling of CuCo2O4 at 60 mA g−1 in the voltage range 0.005–3.0 V versus Li metal exhibits reversible cycling performance between 2 and 50 cycles with a small capacity fading of 2 mAh g−1 per cycle. Good rate capability is also found in the range 0.04–0.94C. Typical discharge and charge capacity values at the 20th cycle are 755(±10) mAh g−1 (∼6.9 mol of Li per mole of CuCo2O4) and 745(±10) mAh g−1 (∼6.8 mol of Li), respectively at a current of 60 mA g−1. The average discharge and charge potentials are ∼1.2 and ∼2.1 V, respectively. The underlying reaction mechanism is the redox reaction: Co ↔ CoO ↔ Co3O4 and Cu ↔ CuO aided by Li2O, after initial reaction with Li. The galvanostatic cycling studies are complemented by cyclic voltammetry (CV), ex situ TEM and SAED. The Li-cycling behaviour of nano-CuCo2O4 compares well with that of iso-structural nano-Co3O4 as reported in the literature.  相似文献   

17.
The composite LaNix/Ni–S–Co film with considerable stability and high HER activity (η150 = 70 mV, 353 K) was obtained by molten salt electrolysis combined with aquatic electrodeposition. LaNix film was prepared by galvanostatic electrolysis at 100 mA cm−2 under 1273 K. The results showed that the La3+ ions could be reduced on the nickel cathode and the LaNix film could form, i.e. La3+ + 3e + xNi = LaNix (x = 5 or 3) at ca. −0.6 V, which is much lower than that of the decomposition potential of lanthanum, due to the strong depolarization effect of nickel. Furthermore, compared with the traditional amorphous Ni–S film, the composite LaNix/Ni–S–Co film could absorb large amount of H atoms, which would be oxidized and avoid the dissolution of the Ni–S–Co film under the state of open-circuit effectively and increase the HER activity.  相似文献   

18.
The relationship between hydrogen generation and the age of culture was investigated under fed-batch growth conditions. The specific growth rate (μe) was determined during the log phase of the growth curve and the μeMax was 0.02643 h−1. Boltzmann's sigmoidal regression model was used to determine the specific rate of hydrogen evolution (μH): the maximum was 0.04440 h−1. At low irradiance (36–75 W m−2), an inverse relationship was found between μH and I; after increasing the irradiance further, μH reached a plateau (0.00916 h−1). The maximum reactor yield of cumulative hydrogen (4.5 l) was obtained at an irradiance of 320 W m−2, but the highest hydrogen evolution rate (17.217 ml h−1) was achieved at 500 W m−2. The light conversion efficiency reached its maximum (6.91%) at the lowest irradiance investigated (36 W m−2); when the irradiance increased further, it decreased progressively down to 0.36%.  相似文献   

19.
This paper quantifies the contribution of Portuguese energy policies for total and marginal abatement costs (MAC) for CO2 emissions for 2020. The TIMES_PT optimisation model was used to derive MAC curves from a set of policy scenarios including one or more of the following policies: ban on nuclear power; ban on new coal power plants without carbon sequestration and storage; incentives to natural gas power plants; and a cap on biomass use. The different MAC shows the policies’ effects in the potential for CO2 abatement. In 2020, in the most encompassing policy scenario, with all current and planned policies, is possible to abate only up to +35% of 1990 emissions at a cost below 23 € t/CO2. In the more flexible policy scenarios, it is possible to abate up to −10% of 1990 emissions below the same cost. The total energy system costs are 10–13% higher if all policies are implemented—76 to 101 B€—roughly the equivalent to 2.01–2.65% of the 2005 GDP. Thus, from a CO2 emission mitigation perspective, the existing policies introduce significant inefficiencies, possibly related to other policy goals. The ban on nuclear power is the instrument that has the most significant effect in MAC.  相似文献   

20.
Electrochemically formed spinel-lithium manganese oxides were synthesized from manganese hydroxides prepared by a cathodic electrochemical precipitation from various concentrations of manganese nitrate solutions. Two types of manganese hydroxides were formed from diluted and concentrated Mn(NO3)2 aqueous solutions. Uniform and equi-sized disk shaped Mn(OH)2 crystals of 0.2–5 μm in diameter were obtained on a Pt substrate after the electrochemical precipitation from lower concentration of ranging from 2 mmol dm−3 to 2 mol dm−3 Mn(NO3)2 aq., while the grass blade-like precipitate which is ascribed to manganese hydroxide with 20–80 μm long and 1–5 μm wide were formed from concentrated Mn(NO3)2 aq.  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