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 15 毫秒
1.
The ternary [Li+][MPPip+][NTf2] ionic liquid, obtained by dissolution of solid lithium bis(trifluoromethanesulfonyl)imide (LiNTf2) in liquid N-methyl-N-propylpiperidinium bis(trifluoromethanesulfonyl)imide (MPPipNTf2), was used as an electrolyte, and stable at the lithium or graphite-lithium anodes. The graphite-lithium (C6Li) anode showed good cyclability and Coulombic efficiency in the presence of a molecular additive (10 wt.% of vinylene carbonate, VC) to the ionic liquid. The electrode showed ca. 90% of its initial discharge capacity after 100 cycles. The addition of ethylene carbonate (EC) does not improve the cyclability of the anode to the same degree as that observed in the case of vinylene carbonate.  相似文献   

2.
Microscopic magnetism of the electrochemically Li-deintercaleted LixCoO2 powders has been investigated by muon-spin rotation and relaxation (μ+SR) spectroscopy in the temperature (T) range between 10 and 300 K. Weak transverse-field μ+SR measurements indicate that localized moments appear in LiCoO2 below 60 K, while both Li0.53CoO2 and Li0.04CoO2 are paramagnetic even at 10 K. Zero-field μ+SR measurements for the samples with x = 0.53 and 0.04 show that the field distribution width (Δ) due to randomly oriented nuclear magnetic moments of 7Li and 59Co decreases monotonically with increasing T up to 250 K, and then it decreases steeper (increasing slope (dΔ/dT)) above 250 K. Because the muon hopping rate (ν) is almost T independent for Li0.53CoO2 below 300 K, the decrease in Δ suggests that the time scale of Li+ diffusion in LixCoO2 is within a microsecond scale.  相似文献   

3.
Lithium metal dendrite growth in Li/poly (ethylene oxide)-lithium bis (trifluoromethanesulfonyl) imide (PEO18LiTFSI), nano-silica, and N-methyl-N-propylpiperidinium bis(trifluoromethanesulfonyl)imide (PP13TFSI) composite solid polymer electrolyte/Li was investigated by direct in situ observation. The dendrite onset time decreased with increasing current density and deviated from Sand's law in the current density range of 0.1-0.5 mA cm−2 at 60 °C. Lithium dendrite formation was not observed until 46 h of polarization at 0.5 mA cm−2 and 60 °C, which is a significant improvement compared to that observed in Li/(PEO18LiTFSI)/Li, where the dendrite formation was observed after 15 h polarization at 0.5 mA cm−2 and 60 °C. The suppression of dendrite formation could be explained by the electrical conductivity enhancement and decrease of the interface resistance between Li and the polymer electrolyte by the introduction of both nano-SiO2 and PP13TFSI into PEO18LiTFSI. The electrical conductivity of 4.96 × 10−4 S cm−1 at 60 °C was enhanced to 7.6 × 10−4 S cm−1, and the interface resistance of Li/PEO18LiTFSI/Li of 248 Ω cm2 was decreased to 74 Ω cm2 by the addition of both nano-SiO2 and PP13TFSI into PEO18LiTFSI.  相似文献   

4.
We investigated the effect of CO2 on layered Li1+zNi1−xyCoxMyO2 (M = Al, Mn) cathode materials for lithium ion batteries which were prepared by solid-state reactions. Li1+zNi(1−x)/2CoxMn(1−x)/2O2 (Ni/Mn mole ratio = 1) singularly exhibited high storage stability. On the other hand, Li1+zNi0.80Co0.15Al0.05O2 samples were very unstable due to CO2 absorption. XPS and XRD measurements showed the reduction of Ni3+ to Ni2+ and the formation of Li2CO3 for Li1+zNi0.80Co0.15Al0.05O2 samples after CO2 exposure. SEM images also indicated that the surfaces of CO2-treated samples were covered with passivation films, which may contain Li2CO3. The relationship between CO2-exposure time and CO32− content suggests that there are two steps in the carbonation reactions; the first step occurs with the excess Li components, Li2O for example, and the second with LiNi0.80Co0.15Al0.05O2 itself. It is well consistent with the fact that the discharge capacity was not decreased and the capacity retention was improved until the excess lithium is consumed and then fast deterioration occurred.  相似文献   

5.
6.
A new series visible-light driven photocatalysts (CuIn)xCd2(1x)S2 was successfully synthesized by a simple and facile, low-temperature hydrothermal method. The synthesized materials were characterized by means of X-ray diffraction (XRD), scanning electron microscopy (SEM), Brunauer–Emmett–Teller (BET) surface area measurement, X-ray photoelectron spectroscopy (XPS) and ultraviolet-visible spectroscopy (UV–Vis DRS). The results show that the morphology of the photocatalysts changes with the increase of x from 0.01 to 0.3 and their band gap can be correspondingly tuned from 2.37 eV to 2.30 eV. The (CuIn)xCd2(1−x)S2 nanocomposite show highly photocatalytic activities for H2 evolution from aqueous solutions containing sacrificial reagents, SO32− and S2− under visible light. Substantially, (CuIn)0.05Cd1.9S2 with the band gap of 2.36 eV exhibits the highest photocatalytic activity even without a Pt cocatalyst (649.9 μmol/(g h)). Theoretical calculations about electronic property of the (CuIn)xCd2(1−x)S2 indicate that Cu 3d and In 5s5p states should be responsible for the photocatalytic activity. Moreover, the deposition of Pt on the doping sample results in a substantial improvement in H2 evolution than the Pt-loaded pure CdS and the amount of H2 produced (2456 μmol/(g h)) in the Pt-loaded doping system is much higher than that of the latter (40.2 μmol/(g h)). The (CuIn)0.05Cd1.9S2 nanocomposite can keep the activity for a long time due to its stability in the photocatalytic process. Therefore, the doping of CuInS2 not only facilitates the photocatalytic activity of CdS for H2 evolution, but also improves its stability in photocatalytic process.  相似文献   

7.
Prospective positive-electrode (cathode) materials for a lithium secondary battery, viz., Li[Li0.2Ni0.2−x/2Mn0.6−x/2Crx]O2 (x = 0, 0.02, 0.04, 0.06, 0.08), were synthesized using a solid-state pyrolysis method. The structural and electrochemical properties were examined by means of X-ray diffraction, cyclic voltammetry, SEM and charge–discharge tests. The results demonstrated that the powders maintain the α-NaFeO2-type layered structure regardless of the chromium content in the range x ≤ 0.08. The Cr doping of x = 0.04 showed improved capacity and rate capability comparing to undoped Li[Li0.2Ni0.2Mn0.6]O2. ac impedance measurement showed that Cr-doped electrode has the lower impedance value during cycling. It is considered that the higher capacity and superior rate capability of Cr-doping samples would be ascribed to the reduced resistance of the electrode during cycling.  相似文献   

8.
The decomposition of crystalline magnesium borohydride upon heating was studied using thermal desorption, calorimetry, in situ X-ray diffraction, and solid state NMR. Hydrogen release from Mg(BH4)2 occurs in at least four steps via formation of several polyborane intermediate species and includes an exothermic reaction yielding crystalline MgH2 as an intermediate. The decomposition products may be only partially recharged after the very first step and also via hydrogenation of Mg metal. The intermediate formation of amorphous MgB12H12, was confirmed by 11B NMR. A four-stage pathway for the thermal decomposition of Mg(BH4)2 is proposed.  相似文献   

9.
An adjustment of a conduction band offset (CBO) of a window/absorber heterointerface is important for high efficiency Cu(In,Ga)Se2 (CIGS) solar cells. In this study, the heterointerface recombination was characterized by the reduction of the thickness of a CdS layer and the adjustment of a CBO value by a Zn1−xMgxO (ZMO) layer. In ZnO/CdS/CIGS solar cells, open-circuit voltage (Voc) and shunt resistance (Rsh) decreased with reducing the CdS thickness. In constant, significant reductions of Voc and Rsh were not observed in ZMO/CdS/CIGS solar cells. With decreasing the CdS thickness, the CBO of (ZnO or ZMO)/CIGS become dominant for recombination. Also, the dominant mechanisms of recombination of the CIGS solar cells are discussed by the estimation of an activation energy obtained from temperature-dependent current-voltage measurements.  相似文献   

10.
In this paper, TiO2 nanotube (TNT) arrays and (Cu2Sn)x/3Zn1−xS (x = 0.75, 0.24, and 0.09) nanocrystals were prepared, and TiO2 nanotubes (TNTs) were sensitized with (Cu2Sn)x/3Zn1−xS nanocrystals. Compared with the plain TNTs, (Cu2Sn)x/3Zn1−xS sensitized TNTs present an enhanced photoelectrochemical response. The photocurrent enhancement was characterized with the photocurrent ratio of (Cu2Sn)x/3Zn1−xS sensitized TNTs to plain TNTs, which were 1.50, 1.63, and 2.13 for (Cu2Sn)x/3Zn1−xS with the composition of x = 0.75, 0.24, and 0.09, respectively. To understand this phenomenon, the energy band alignments of TNTs and (Cu2Sn)x/3Zn1−xS were investigated, based on which the conduction band offsets (CBO) between TNTs and (Cu2Sn)x/3Zn1−xS were determined, which were 0.31, 0.47, and 0.63 eV for (Cu2Sn)x/3Zn1−xS with the composition of x = 0.75, 0.24, and 0.09, respectively. These results display that the photocurrent enhancement becomes large with the increase of CBO, which indicates that the enhanced photoelectrochemical response is mainly due to the energy level matching between TNTs and (Cu2Sn)x/3Zn1−xS, and the variation of enhancement is resulting from the change of CBO.  相似文献   

11.
Quantitative phase analysis of Cu(In1−xGax)Se2 (CIGS) thin film grown over Mo coated soda lime glass substrates was studied by Rietveld refinement process using room temperature X-ray data at θ-2θ mode. Films were found to contain both stoichiometric Cu(In1−xGax)Se2 and defect related Cu(In1−xGax)3Se5 phases. Best fitting was obtained using crystal structure with space group I-42d for Cu(In1−xGax)Se2 and I-42m for Cu(In1−xGax)3Se5 phase. The effects of Ga/III (=Ga/In+Ga=x) ratio and Se flux during growth over the formation of Cu(In1−xGax)3Se5 defect phase in CIGS was studied and the correlation between quantity of Cu(In1−xGax)3Se5 phase and solar cell performance is discussed.  相似文献   

12.
The composites of (NaBH4+2Mg(OH)2) and (LiBH4+2Mg(OH)2) without and with nanometric Ni (n-Ni) added as a potential catalyst were synthesized by high energy ball milling. The ball milled NaBH4-based composite desorbs hydrogen in one exothermic reaction in contrast to its LiBH4-based counterpart which dehydrogenates in two reactions: an exothermic and endothermic. The NaBH4-based composite starts desorbing hydrogen at 240 °C. Its ball milled LiBH4-based counterpart starts desorbing at 200 °C. The latter initially desorbs hydrogen rapidly but then the rate of desorption suddenly decelerates. The estimated apparent activation energy for the NaBH4-based composite without and with n-Ni is equal to 152 ± 2.2 and 157 ± 0.9 kJ/mol, respectively. In contrast, the apparent activation energy for the initial rapid dehydrogenation for the LiBH4-based composite is very low being equal to 47 ± 2 and 38 ± 9 kJ/mol for the composite without and with the n-Ni additive, respectively. XRD phase studies after volumetric isothermal dehydrogenation tests show the presence of NaBO2 and MgO for the NaBH4-based composite. For the LiBH4-based composite phases such as MgO, Li3BO3, MgB2, MgB6 are the products of the first exothermic reaction which has a theoretical H2 capacity of 8.1 wt.%. However, for reasons which are not quite clear, the first reaction never goes to full completion even at 300 °C desorbing ∼4.5 wt.% H2 at this temperature. The products of the second endothermic reaction for the LiBH4-based composite are MgO, MgB6, B and LiMgBO3 and the reaction has a theoretical H2 capacity of 2.26 wt.%. The effect of the addition of 5 wt.% nanometric Ni on the dehydrogenation behavior of both the NaBH4-and LiBH4-based composites is rather negligible. The n-Ni additive may not be the optimal catalyst for these hydride composite systems although more tests are required since only one n-Ni content was examined.  相似文献   

13.
Electrolytes containing LiB(C2O4)2 (LiBOB) salts are of increasing interest for lithium-ion cells for several reasons that include their ability to form a stable solid electrolyte interphase on graphite electrodes. However, cells containing these electrolytes often show inconsistent performance because of impurities in the LiBOB salt. In this work we compare cycling and impedance data from cells containing electrolytes with LiBOB that was obtained commercially and LiBOB purified by a rigorous recrystallization procedure. We relate the difference in performance to a lithium oxalate impurity that may be a residual from the salt manufacturing process. We also examine the reaction of LiBOB with water to determine the effect of salt storage in high-humidity environments. Although LiBOB electrolytes containing trace amounts (∼100 ppm) of moisture appear relatively stable, higher moisture contents (∼1 wt%) lead to observable salt decomposition resulting in the generation of B(C2O4)(OH) and LiB(C2O4)(OH)2 compounds that do not dissolve in typical carbonate solutions and impair lithium-ion cell performance.  相似文献   

14.
A new type of Li1−xFe0.8Ni0.2O2–LixMnO2 (Mn/(Fe + Ni + Mn) = 0.8) material was synthesized at 350 °C in air atmosphere using a solid-state reaction. The material had an XRD pattern that closely resembled that of the original Li1−xFeO2–LixMnO2 (Mn/(Fe + Mn) = 0.8) with much reduced impurity peaks. The Li/Li1−xFe0.8Ni0.2O2–LixMnO2 cell showed a high initial discharge capacity above 192 mAh g−1, which was higher than that of the parent Li/Li1−xFeO2–LixMnO2 (186 mAh g−1). We expected that the increase of initial discharge capacity and the change of shape of discharge curve for the Li/Li1−xFe0.8Ni0.2O2–LixMnO2 cell is the result from the redox reaction from Ni2+ to Ni3+ during charge/discharge process. This cell exhibited not only a typical voltage plateau in the 2.8 V region, but also an excellent cycle retention rate (96%) up to 45 cycles.  相似文献   

15.
Computational and experimental work directed at exploring the electrochemical properties of tetrahedrally coordinated Mn in the +5 oxidation state is presented. Specific capacities of nearly 700 mAh g−1 are predicted for the redox processes of LixMnO4 complexes based on two two-phase reactions. One is topotactic extraction of Li from Li3MnO4 to form LiMnO4 and the second is topotactic insertion of Li into Li3MnO4 to form Li5MnO4. In the experiments, it is found that the redox behavior of Li3MnO4 is complicated by disproportionation of Mn5+ in solution to form Mn4+ and Mn7+ and by other irreversible processes; although an initial capacity of about 275 mAh g−1 in lithium cells was achieved. Strategies based on structural considerations to improve the electrochemical properties of MnO4n complexes are given.  相似文献   

16.
The electrochemical reactivity of the layered titanium hydrogeno phosphate Ti(HPO4)2·H2O versus lithium has been studied. Lithium intercalation occurs at ∼2.5 V with low polarization, leading to a new lithiated Ti(III) phase, LiTi(HPO4)2·H2O. Ti(HPO4)2·H2O exhibits a reversible capacity of 80 mAh g−1 in the voltage window 1.8–3.5 V at C/10 rate. The stable reversible capacity reveals that the presence of H2O lattice is not affecting the electrochemical reaction. The reversibility of the reaction is demonstrated by extracting lithium from LiTi(HPO4)2·H2O and the host structure is intact. The electrochemical behaviour of dehydrated phases Ti(HPO4)2 and TiP2O7 has also been investigated.  相似文献   

17.
The structure, vibration, and electronic structure of H2 molecule adsorbed on (ZrO2)n (n = 1-6) clusters were investigated with density functional theory. We found that H2 is easily absorbed on the top Zr atoms of (ZrO2)n (n = 1-6) clusters. The Zr5O10H2 cluster has the lowest binding energies in the ZrnO2nH2 (n = 1-6) clusters. By analyzing vibrational frequency and Mulliken charge, the H-O and Zr-H bonds were found to be formed in different sized ZrnO2nH2 clusters. The dissociation mechanism of H2 shows that the charge transfers from (ZrO2)n cluster to H2 due to the important role of the orbital hybridization between the cluster and H2 molecule. With increasing the number of H2 molecule adsorbed on (ZrO2)n clusters, the adsorption favors to the sites with low coordinate number, and these adsorption modes present a symmetrical tendency.  相似文献   

18.
The temperature dependence of open-circuit voltage (Voc), short-circuit current (Isc), fill factor (FF), and relative efficiency of monograin Cu2ZnSn(SexS1−x)4 solar cell was measured. The light intensity was varied from 2.2 to 100 mW/cm2 and temperatures were in the range of = 175-300 K. With a light intensity of 100 mW/cm2dVoc/dT was determined to be −1.91 mV/K and the dominating recombination process at temperatures close to room temperature was found to be related to the recombination in the space-charge region. The solar cell relative efficiency decreases with temperature by 0.013%/K. Our results show that the diode ideality factor n does not show remarkable temperature dependence and slightly increases from n = 1.85 to n = 2.05 in the temperature range between 175 and 300 K.  相似文献   

19.
A stable quasi-solid-state dye-sensitized solar cell (DSC) with a novel amphiphilic polymer gel electrolyte (APGE) based on poly(lactic acid-co-glycolic acid) (PLGA) is fabricated. The APGE could be readily prepared by a simple method at low temperature of 50 °C and exhibits a quasi-solid property, high conductivity, and long-term stability. The 20 and 40 wt% APGE-based DSCs show high photovoltaic conversion efficiency of 7.5 and 7.4%, respectively, under AM 1.5 simulated sunlight, which is comparable to the liquid electrolyte-based DSC with the efficiency of 7.6%. The 40 wt% APGE-based DSC maintains 95% of the initial performance after 60 days in practical conditions. It is also noteworthy that the APGE endows with higher short-circuit current density than the liquid electrolyte. Different natures of the APGE from the typical polymer gel electrolytes have been elucidated by the I-V measurements, electrochemical impedance spectroscopy, electrophoretic measurements, and transmission electron microscopy.  相似文献   

20.
Molecular hydrogen (H2) production by Escherichia coli was studied during mixed carbon sources (glucose and glycerol) fermentation at pH 7.5 and pH 5.5. H2 production rate (VH2) by bacterial cells grown on mixed carbon was assayed with either adding glucose (glucose assay) or glycerol (glycerol assay) and compared with the cells grown on sole carbon (glucose or glycerol only) and appropriately assayed. Wild type cells grown on mixed carbon, in the assays with adding glucose, produced H2 at pH 7.5 with the same level as in the cells grown on glucose only. At pH 7.5 VH2 in fhlA single and fhlA hyfG double mutants decreased ∼6.5 and ∼7.9 fold, respectively. In wild type cells grown on mixed carbon VH2 at pH 5.5 was lowered ∼2 fold, compared to the cells grown on glucose only. But in hyfG and hybC single mutants VH2 was decreased ∼2 and ∼1.6 fold, respectively. However, at pH 7.5, in the assays with glycerol, VH2 was low, when compared to the cells grown on glycerol only. At pH 5.5 in the assays with glycerol VH2 was absent. Moreover, VH2 in wild type cells was inhibited by 0.3 mM N,N-dicyclohexylcarbodiimide (DCCD), an inhibitor of the F0F1-ATPase, in a pH dependent manner. At pH 7.5 in wild type cells VH2 was decreased ∼3 fold but at pH 5.5 the inhibition was ∼1.7 fold. At both pHs in fhlA mutant VH2 was totally inhibited by DCCD. Taken together, the results obtained indicate that at pH 7.5, in the presence of glucose, glycerol can also be fermented. They point out that Hyd-4 mainly and Hyd-2 to some extent contribute in H2 production by E. coli during mixed carbon fermentation at pH 5.5 whereas Hyd-1 is only responsible for H2 oxidation.  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