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.
In CdS/CdTe solar cells it is necessary to determine the efficiency limitations related with the intermixing at the interface between the CdS window layer and CdTe absorber layer. So understanding the properties of the solid solution (CdSxTe1?x on the CdTe side which is CdTe rich and CdS1?yTey on the CdS side which is CdS rich) that is always formed in this region is essential. We produced thin films of CdS1?yTey solid solution-which is CdS rich – by first producing CdS:In thin films on glass substrates by the spray pyrolysis technique and then annealing the films in nitrogen atmosphere at 400 °C in the presence of Te vapor. We are the first who produce this solid solution by this simple and low cost method. The composition and morphology of the films were determined by energy dispersive X-ray detection (EDAX) measurements and scanning electron microscopy (SEM) observations respectively. Eight values of y in the range 0 ? y ? 0.2845 were obtained. The transmittance was measured and used to investigate the optical bandgap energy by using the second derivative of the absorbance. It is found that the films show a single hexagonal phase for y ? 0.0852 and then a mixed (hexagonal and cubic) phase for 0.0997 ? y ? 0.2845. Bandgap energies in the range 2.259 ? Eg ? 2.528 eV were obtained. Urbach tailing in the bandgap was also investigated.  相似文献   

2.
《Journal of power sources》2006,159(1):249-253
The chemical and structural stabilities of various layered Li1−xNi1−yzMnyCozO2 cathodes are compared by characterizing the samples obtained by chemically extracting lithium from the parent Li1−xNi1−yzMnyCozO2 with NO2BF4 in an acetonitrile medium. The nickel- and manganese-rich compositions such as Li1−xNi1/3Mn1/3Co1/3O2 and Li1−xNi0.5Mn0.5O2 exhibit better chemical stability than the LiCoO2 cathode. While the chemically delithiated Li1−xCoO2 tends to form a P3 type phase for (1  x) < 0.5, Li1−xNi0.5Mn0.5O2 maintains the original O3 type phase for the entire 0  (1  x)  1 and Li1−xNi1/3Mn1/3Co1/3O2 forms an O1 type phase for (1  x) < 0.23. The variations in the type of phases formed are explained on the basis of the differences in the chemical lithium extraction rate caused by the differences in the degree of cation disorder and electrostatic repulsions. Additionally, the observed rate capability of the Li1−xNi1−yzMnyCozO2 cathodes bears a clear relationship to cation disorder and lithium extraction rate.  相似文献   

3.
Powder of nanosized particles of Ru-based (Rux, RuxSey and RuxFeySez) clusters were prepared as catalysts for oxygen reduction in 0.5 M H2SO4 and for fuel cells prepared by pyrolysis in organic solvent. These electrocatalysts show a high uniformity of agglomerated nanometric particles. The reaction kinetics were studied using rotating disk electrodes and an enhanced catalytic activity for the powders containing selenium and iron was observed. The Ru-based electrocatalysts were used as the cathode in a single prototype PEM fuel cell, which was prepared by spray deposition of the catalyst on the surface of Nafion® 117 membranes. The electrochemical performance of each single fuel cell was compared to that of a platinum/platinum conventional membrane electrode assembly (MEA), using hydrogen and oxygen feed streams. A maximum power density of 140 mW cm−2, at 80 °C with 460 mA cm−2 was obtained for the RuxFeySez catalysts; approximately 55% lower power density than that obtained with platinum.  相似文献   

4.
The preparation, structural characterization and study of the electrochemical behavior in lithium cells of Co1−2yFeyNiySb3 (0.125<y≤0.5) compounds is described. The refinement of X-ray diffraction (XRD) patterns evidenced the skutterudite-type structure for all compositions, with an increasing partial filling of 2a sites. The first discharge of lithium anode cells shows two plateaus. The plateau at higher potential, which occurs at ca. 0.8 V and extends from ca. 150 to 250 Ah/kg depending on “y”, is mainly assigned to side reactions of lithium with the electrolyte, as confirmed by 7Li NMR. The second plateau occurs at 0.5–0.6 V and is assigned to Li–Sb alloys formation. The charge of the cell shows that only the second step is reversible. Further discharge/charge cycles present a plateau at 0.8 V in the discharge and at 1.05 V in the charge, which agree well with the Li–Sb alloying/de-alloying process. Extended cycling results on a loss of capacity, to stabilize over 100 Ah/kg after 20 cycles.  相似文献   

5.
《Journal of power sources》2006,159(2):1310-1315
We report the epitaxial growth of the LiNi1−yMyO2 films (M = Co, Co–Al) on heated nickel foil using pulsed laser deposition in oxygen environment from lithium-rich targets. The structure and morphology was characterized by X-ray diffractometry, electron scanning microscopy and Raman spectroscopy. Data reveal that the formation of oriented films is dependent on two important parameters: the substrate temperature and the gas pressure during ablation. The charge–discharge process conducted in Li-microcells demonstrates that effective high specific capacities can be obtained with films 1.35 μm thick. Stable capacities of 83 and 92 μAh cm−2 μm are available in the potential range 4.2–2.5 V for LiNi0.8Co0.2O2 and LiNi0.8Co0.15Al0.05O2 films, respectively. The self-diffusion coefficient of Li ions determined from galvanostatic intermittent titration experiments is found to be 4 × 10−12 cm2 s−1.  相似文献   

6.
《Journal of power sources》2003,115(1):119-124
Insertion of lithium and sodium into phosphate (MoO2)2P2O7 was investigated electrochemically to determine the usefulness as a possible cathode for ion-transfer secondary batteries. Specific charges of up to 250 mA h g−1 were obtained for A/(MoO2)2P2O7 (A: Li, Na) cells with liquid organic electrolytes in the first reduction half-cycle at room temperature. Intercalation processes under constant current densities of 0.2 mA cm−2 were reversible within the range of composition 0.85<x<4.0 for lithium and 0.5<x<3.1 for sodium in Ax(MoO2)2P2O7 (A: Li, Na), respectively. Structural changes induced by lithium or sodium intercalation were followed by ex situ X-ray diffraction measurements, and the phase change from the crystal to the amorphous was observed in both cases.  相似文献   

7.
《Journal of power sources》2005,141(1):198-203
The relationship between the structure-specific capacitance (F g−1) of a composite electrode consisting of activated coconut-shell carbon and hydrous ruthenium oxide (RuOx(OH)y) has been evaluated by impregnating various amounts of RuOx(OH)y into activated carbon that is specially prepared with optimum pore-size distribution. The composite electrode shows an enhanced specific capacitance of 250 F g−1 in 1 M H2SO4 with 9 wt.% ruthenium incorporated. Chemical and structural characterization of the composites reveals a homogeneous distribution of amorphous RuOx(OH)y throughout the porous network of the activated carbon. Electrochemical characterization indicates an almost linear dependence of capacitance on the amount of ruthenium owing to its pseudocapacitive nature.  相似文献   

8.
《Journal of power sources》2006,153(2):402-404
Thermal reactions of a lithiated graphite anode in 1 M LiPF6-ethylene carbonate (EC)/dimethyl carbonate (DMC) (50:50 vol.%) in the temperature range 40–320 °C were investigated by TG–MS analysis. Studies by TG–MS during thermal reactions detected a small exothermic peak around 140 °C due to CO2 (m/z = 44) evolution, which suggests partial destruction of the SEI formed on the graphite and/or decomposition of the electrolyte through the SEI. In addition, the main exothermic reaction above 280 °C, which is associated with simultaneous evolution of C2H4O (m/z = 44), is caused by direct reaction of the lithiated graphite with solvent.  相似文献   

9.
《Journal of power sources》2004,133(2):268-271
Following the route of synthesis of β-MoO3 through soft chemistry methods a new amorphous material with composition MoO3·2H2O has been detected. The hydrated molybdenum oxide showed the capacity for electrochemical lithium insertion. The maximum amount of lithium incorporated in this material (∼3.3 Li/Mo) leads to a specific capacity of 490 Ah kg−1. The charge–discharge curve showed a good reversibility in the potential range from 3.2 to 1.1 V versus Li+/Li0 where the cell voltage decreased monotonously as a function of the degree of lithium inserted. The electrochemical features of amorphous MoO3·2H2O suggest that it can be considered as a possible cathode candidate in rechargeable lithium batteries.  相似文献   

10.
《Journal of power sources》2006,162(1):667-672
The crystal chemistry and electrochemical performance of the layered LiNi0.5−yCo0.5−yMn2yO2 and LiCo0.5−yMn0.5−yNi2yO2 oxide cathodes for 0  2y  1 have been investigated. Li2MnO3 impurity phase is observed for Mn-rich compositions with 2y > 0.6 in LiNi0.5−yCo0.5−yMn2yO2 and 2y < 0.2 in LiCo0.5−yMn0.5−yNi2yO2. Additionally, the Ni-rich compositions encounter a volatilization of lithium at the high synthesis temperature of 900 °C. Compositions around 2y = 0.33 are found to be optimum with respect to maximizing the capacity values and retention. The rate capabilities are found to bear a strong relationship to the cation disorder in the layered lattice. Moreover, the evolution of the X-ray diffraction patterns on chemically extracting lithium has revealed the presence of Li2MnO3 phase in addition to the layered phase for the composition LiNi0.25Co0.25Mn0.5O2 with an oxidation state of manganese close to 4+, which results in a large anodic peak at around 4.5 V due to the extraction of both lithium and oxygen.  相似文献   

11.
《Journal of power sources》2006,162(2):1312-1321
Lithium insertion and extraction in to/from the oxyfluorides TiOF2 and NbO2F is investigated by galvanostatic cycling, cyclic voltammetry and impedance spectroscopy in cells using Li-metal as a counter electrode at ambient temperature. The host compounds are prepared by low-temperature reaction and characterized by powder X-ray diffraction (XRD), Rietveld refinement and Brunauer, Emmett and Teller (BET) surface area. Crystal structure destruction occurs during the first-discharge reaction with Li at voltages below 0.8–0.9 V for LixTiOF2 as shown by ex situ XRD and at ≤1.4 V for LixNbO2F to form amorphous composites, ‘LixTi/NbOy–LiF’. Galvanostatic discharge–charge cycling of ‘LixTiOy’ in the range 0.005–3.0 V at a current density of 65 mA g−1 gives a capacity of 400 (±5) mAh g−1 during 5–100 cycles with no noticeable capacity fading. This value corresponds to 1.52 mol of recycleable Li/Ti. The coulombic efficiency (η) is >98%. Results on ‘LixNbOy’ show good reversibility of the electrode and a η >98% is achieved only after 10 cycles (range 0.005–3.0 V and at 30 mA g−1) and a capacity of 180 (±5) mAh g−1 (0.97 mol of Li/Nb) was stable up to 40 cycles. In both ‘LixTiOy’ and ‘LixNbOy’, the average discharge and charge voltages are 1.2–1.4 and 1.7–1.8 V, respectively. The impedance spectral data measured during the first cycle and after selected numbers of cycles are fitted to an equivalent circuit and the roles played by the relevant parameters as a function of cycle number are discussed.  相似文献   

12.
《Journal of power sources》2006,158(1):641-645
Stabilized lithium nickelate is receiving increased attention as a low-cost alternative to the LiCoO2 cathode now used in rechargeable lithium batteries. Layered LiNi1−xyMxMyO2 samples (Mx = Al3+ and My = Mg2+, where x = 0.05, 0.10 and y = 0.02, 0.05) are prepared by the refluxing method using acetic acid at 750 °C under an oxygen stream, and are subsequently subjected to powder X-ray diffraction analysis and coin-cell tests. The co-doped LiNi1−xyAlxMgyO2 samples show good structural stability and electrochemical performance. The LiNiAl0.05Mg0.05O2, cathode material exhibits a reversible capacity of 180 mA h g−1 after extended cycling. These results suggest that the threshold concentration for aluminum and magnesium substitution is of the order of 5%. The co-substitution of magnesium and aluminium into lithium nickelate is considered to yield a promising cathode material.  相似文献   

13.
In-situ ultra-thin porous poly(vinylidene fluoride-co-hexafluoropropylene) P(VDF–HFP) membranes were prepared by a phase inversion method on TiO2 electrodes coated with Ru N-719 dye. These membranes were then soaked in the organic liquid electrolyte to form the in-situ ultra-thin porous P(VDF–HFP) membrane electrolytes. Dye-sensitized solar cell (DSC) using the membrane electrolyte exhibited an open-circuit voltage (Voc) of 0.751 V, a short-circuit current (Jsc) of 16.260 mA cm?2 and a fill factor (FF) of 0.684 under an incident light intensity of 1000 W m?2 yielding an energy conversion efficiency (η) of 8.35%. The Voc, FF and η of the solar cell using the membrane electrolyte increased by about 5.8%, 2.2% and 5.7%, respectively, when compared with the corresponding values of a cell using liquid electrolyte. However, the Jsc decreased by about 2.1%.  相似文献   

14.
《Journal of power sources》2006,162(2):1036-1042
To protect the ceria electrolyte from reduction at the anode side, a thin film of yttria-stabilized zirconia (YSZ) is introduced as an electronic blocking layer to anode-supported gadolinia-doped ceria (GDC) electrolyte solid oxide fuel cells (SOFCs). Thin films of YSZ/GDC bilayer electrolyte are deposited onto anode substrates using a simple and cost-effective wet ceramic co-sintering process. A single cell, consisting of a YSZ (∼3 μm)/GDC (∼7 μm) bilayer electrolyte, a La0.8Sr0.2Co0.2Fe0.8O3–GDC composite cathode and a Ni–YSZ cermet anode is tested in humidified hydrogen and air. The cell exhibited an open-circuit voltage (OCV) of 1.05 V at 800 °C, compared with 0.59 V for a single cell with a 10-μm GDC film but without a YSZ film. This indicates that the electronic conduction through the GDC electrolyte is successfully blocked by the deposited YSZ film. In spite of the desirable OCVs, the present YSZ/GDC bilayer electrolyte cell achieved a relatively low peak power density of 678 mW cm−2 at 800 °C. This is attributed to severe mass transport limitations in the thick and low-porosity anode substrate at high current densities.  相似文献   

15.
《Journal of power sources》2006,158(2):1202-1208
Nd0.6Sr0.4Co1−yMnyO3−δ (0  y  1.0) oxides have been investigated as cathode materials for intermediate temperature solid oxide fuel cells (SOFC). The samples form a single-phase solid solution with an orthorhombic perovskite structure and the lattice parameters and volume increase with increasing Mn content y. The degree of oxygen loss at high temperatures and the thermal expansion coefficient (TEC) decrease with increasing y due to a stronger MnO bond. The electrical conductivity decreases with y and the system exhibits a metal to semiconductor transition at around y = 0.2. The electrocatalytic activity and power density measured with single cell SOFC decrease with increasing y due to a decrease in oxygen exchange and mobility as well as charge transfer kinetics, arising from a decrease in the oxide ion vacancy concentration and electrical conductivity.  相似文献   

16.
Anode-supported solid oxide fuel cells (SOFCs) with lanthanum-doped ceria (LDC)/Sr-, Mg-doped LaGaO3 (LSGM) bilayered or LDC/LSGM/LDC trilayered electrolyte films were fabricated with a pure La0.6Sr0.4CoO3 (LSC) cathode. The behaviors of the two electrolytes in cells were investigated by using scanning electron microscopy, impedance spectroscopy and cell performance measurements. The reactions between LSGM and anode material can be suppressed by applying a ca. 15 μm LDC film. Due to the Co diffusion from the LSC cathode to the LSGM electrolyte during high temperature sintering, the electronic conductivity of the LDC electrolyte cannot be completely blocked with an LSGM layer below 50 μm, which leads to open-circuit potentials of these cells of ca. 0.988 V at 800 °C. The electrical conductivities of LDC and LSGM electrolytes in the cells under operation conditions are obtained from the dependence of the cell ohmic resistance on the electrolyte thickness. The electrical conductivity of LDC electrolyte is ca. 0.117 S cm−1 at 800 °C on the bilayered electrolyte cells with a 50 μm LSGM layer. The bilayer electrolyte cells with a 25 μm LDC layer at 800 °C, had a cell ohmic resistance two-stage linear dependence on the LSGM layer thickness, which showed the electrical conductivity of ca. 1.9 S cm−1 for the LSGM layer below 50 μm and 0.22 S cm−1 for the LSGM layer above 100 μm. With a LDC/LSGM/LDC trilayered electrolyte film for the anode-supported cell, an open-circuit potential of 1.043 V was achieved.  相似文献   

17.
《Biomass & bioenergy》2006,30(10):892-896
Anaerobic treatment of solid wastes from potato processing was studied in completely stirred tank reactors (CSTR) at 55 °C. Special attention was paid to the effect of increased organic loading rate (OLR) on the biogas yield in long-term experiments. Both biogas yield and CH4 in the biogas decreased with the increase in OLR. For OLR in the range of 0.8 gl−1 d−1–3.4 gl−1 d−1, biogas yield and CH4 obtained were 0.85 l g−1–0.65 l g−1 and 58%–50%, respectively. Biogas yield y as a function of maximum biogas yield ym, reaction rate constant k and HRT are described on the basis of a mass balance in a CSTR and a first order kinetic. The value of ym can be obtained from curve fitting or a simple batch test and k results from plotting y/(ymy) against 1/OLR from long-term experiments. In the present study values for ym and k were obtained as 0.88 l g−1 and 0.089 d−1, respectively. The simple model equations can apply for dimensioning completely stirred tank reactors (CSTR) digesting organic wastes from food processing industries, animal waste slurries or biogas crops.  相似文献   

18.
The electrochemical performance of a hydrogen sulfide solid oxide fuel cell having the configuration H2S, Pt/(ZrO2)0.92(Y2O3)0.08/Pt, air has been examined at atmospheric pressure and 750–800°C, using both pure and 5% H2S anode feed streams. The performance of the cell is higher when using diluted H2S feed compared with pure H2S feed: current densities up to 100 mA cm−2 and power densities up to 15.4 mW cm−2 have been achieved using diluted H2S gas (5%) at 800°C. However, the platinum anode degrades over time in H2S stream due to the formation of PtS. Electrochemical oxidation of H2S on the Pt anode significantly accelerated its degradation. Polarization and impedance spectroscopy measurements show that at low current density (i) electrochemical reaction is the major cause of polarization in the fuel cell. Ohmic loss due to the resistance of the electrolyte material and the electrical connecting wire is a major part of cell polarization at high i.  相似文献   

19.
《Journal of power sources》2006,161(2):1056-1061
Ni–Cu alloy-based anodes, Ni1−xCux (x = 0, 0.05, 0.2, 0.3)–Ce0.8Sm0.2O1.9 (SDC), were developed for direct utilization of biomass-produced gas in low-temperature solid oxide fuel cells (LT-SOFCs) with thin film Ce0.9Gd0.1O1.95 electrolytes. The alloys were formed by in situ reduction of Ni1−xCuxOy composites synthesized using a glycine-nitrate technique. The electrolyte films were fabricated with a co-pressing and co-firing technique. Electrochemical performance of the Ni1−xCux–SDC anode supported cells was investigated at 600 °C when humidified (3% H2O) biomass-produced gas (BPG) was used as the fuel and stationary air as the oxidant. With Ni–Cu alloys as anodes, carbon deposition was substantially suppressed and electrochemical performance of the cells was sustained for much longer periods of time. For example, the power export of a Ni–SDC supported cell was only 50% of the initial value (200 mW cm−2 at 0.5 V) after 20 min, while Ni0.8Cu0.2–SDC supported cells could maintain 90% of the initial power density (250 mW cm−2 at 0.5 V) over a period of 10 h. The improved performance of the Ni–Cu alloy-based anodes is worth considering in developing SOFCs fueled directly with dilute hydrocarbons such as gases derived from biomass.  相似文献   

20.
Thermodynamic equilibrium of methanol steam reforming (MeOH SR) was studied by Gibbs free minimization for hydrogen production as a function of steam-to-carbon ratio (S/C = 0–10), reforming temperature (25–1000 °C), pressure (0.5–3 atm), and product species. The chemical species considered were methanol, water, hydrogen, carbon dioxide, carbon monoxide, carbon (graphite), methane, ethane, propane, i-butane, n-butane, ethanol, propanol, i-butanol, n-butanol, and dimethyl ether (DME). Coke-formed and coke-free regions were also determined as a function of S/C ratio.Based upon a compound basis set MeOH, CO2, CO, H2 and H2O, complete conversion of MeOH was attained at S/C = 1 when the temperature was higher than 200 °C at atmospheric pressure. The concentration and yield of hydrogen could be achieved at almost 75% on a dry basis and 100%, respectively. From the reforming efficiency, the operating condition was optimized for the temperature range of 100–225 °C, S/C range of 1.5–3, and pressure at 1 atm. The calculation indicated that the reforming condition required from sufficient CO concentration (<10 ppm) for polymer electrolyte fuel cell application is too severe for the existing catalysts (Tr = 50 °C and S/C = 4–5). Only methane and coke thermodynamically coexist with H2O, H2, CO, and CO2, while C2H6, C3H8, i-C4H10, n-C4H10, CH3OH, C2H5OH, C3H7OH, i-C4H9OH, n-C4H9OH, and C2H6O were suppressed at essentially zero. The temperatures for coke-free region decreased with increase in S/C ratios. The impact of pressure was negligible upon the complete conversion of MeOH.  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