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.
《Journal of power sources》2006,155(2):340-352
Thermodynamic analyses of producing a hydrogen-rich fuel-cell feed from the combined processes of dimethyl ether (DME) partial oxidation and steam reforming were investigated as a function of oxygen-to-carbon ratio (0.00–2.80), steam-to-carbon ratio (0.00–4.00), temperature (100 °C–600 °C), pressure (1–5 atm) and product species.Thermodynamically, dimethyl ether processed with air and steam generates hydrogen-rich fuel-cell feeds; however, the hydrogen concentration is less than that for pure DME steam reforming. Results of the thermodynamic processing of dimethyl ether indicate the complete conversion of dimethyl ether to hydrogen, carbon monoxide and carbon dioxide for temperatures greater than 200 °C, oxygen-to-carbon ratios greater than 0.00 and steam-to-carbon ratios greater than 1.25 at atmospheric pressure (P = 1 atm). Increasing the operating pressure has negligible effects on the hydrogen content. Thermodynamically, dimethyl ether can produce concentrations of hydrogen and carbon monoxide of 52% and 2.2%, respectively, at a temperature of 300 °C, and oxygen-to-carbon ratio of 0.40, a pressure of 1 atm and a steam-to-carbon ratio of 1.50. The order of thermodynamically stable products (excluding H2, CO, CO2, DME, NH3 and H2O) in decreasing mole fraction is methane, ethane, isopropyl alcohol, acetone, n-propanol, ethylene, ethanol and methyl-ethyl ether; trace amounts of formaldehyde, formic acid and methanol are observed.Ammonia and hydrogen cyanide are also thermodynamically favored products. Ammonia is favored at low temperatures in the range of oxygen-to-carbon ratios of 0.40–2.50 regardless of the steam-to-carbon ratio employed. The maximum ammonia content (i.e., 40%) occurs at an oxygen-to-carbon ratio of 0.40, a steam-to-carbon ratio of 1.00 and a temperature of 100 °C. Hydrogen cyanide is favored at high temperatures and low oxygen-to-carbon ratios with a maximum of 3.18% occurring at an oxygen-to-carbon ratio of 0.40 and a steam-to-carbon ratio of 0.00 in the temperature range of 400 °C–500 °C. Increasing the system pressure shifts the equilibrium toward ammonia and hydrogen cyanide.  相似文献   

2.
《Journal of power sources》2006,158(1):485-496
This paper describes a study of steam reforming of methane using unsupported nickel powder catalysts. The reaction yields were measured and the unsupported nickel powder surface was studied to explore its potential as a catalyst in internal or external reforming solid oxide fuel cells. The unsupported nickel catalyst used and presented in this paper is a pure micrometric nickel powder with an open filamentary structure, irregular ‘fractal-like’ surface and high external/internal surface ratio. CH4 conversion increases and coke deposition decreases significantly with the decrease of CH4:H2O ratio. At a CH4:H2O ratio of 1:2 thermodynamic equilibrium is achieved, even with methane residence times of only ∼0.5 s. The CH4 conversion is 98 ± 2% at 700 °C and no coke is generated during steam reforming which compares favorably with supported Ni catalyst systems. This ratio was used in further investigations to measure the hydrogen production, the CH4 conversion, the H2 yield and the selectivity of the CO, and CO2 formation. Methane-rich fuel ratios cause significant deviations of the experimental results from the theoretical model, which has been partially correlated to the adsorption of carbon on the surface according to TEM, XPS and elemental analysis. At the fuel: water ratio of 1:2, the unsupported Ni catalyst exhibited high catalytic activity and stability during the steam reforming of methane at low-medium temperature range.  相似文献   

3.
《Journal of power sources》2006,159(2):817-823
The performance of a poly(2,5-benzimidazole) (ABPBI) membrane based high temperature PEM fuel cell in presence of carbon monoxide, at various temperatures is reported here. The ABPBI was synthesized by polymerization of 3,4-diaminobenzoic acid in a polymerization medium containing methanesulfonic acid (CH3SO3H) and phosphorous pentoxide (P2O5). The ABPBI membranes were characterized by fourier transform infrared spectroscopy (FT-IR) and scanning electron microscopy (SEM). A maximum conductivity of 0.026 S cm−1 at 180 °C was obtained for the membrane doped with 1.2 molecules of phosphoric acid (H3PO4) per polymer repeat unit. Fuel cell performance was evaluated using dry hydrogen/oxygen gases and was comparable with that reported in the literature. Performance of a single cell at different temperatures was studied with 0.48 and 1.0 vol.% of CO in the hydrogen fuel. The studies lead to the conclusion that CO poisoning is not a serious problem above 170 °C. Performance of the fuel cell operating at 210 °C is not at all affected by 1.0 vol.% of CO in the hydrogen feed.  相似文献   

4.
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   

5.
《Journal of power sources》2006,157(1):104-113
This paper presents a comprehensive study of hydrogen production from sodium borohydride (NaBH4), which is synthesized from sodium tetraborate (Na2B4O7) decomposition, for proton exchange membrane (PEM) fuel cells. For this purpose, Na2B4O7 decomposition reaction at 450–500 °C under hydrogen atmosphere and NaBH4 decomposition reaction at 25–40 °C under atmospheric pressure are selected as a common temperature range in practice, and the inlet molar quantities of Na2B4O7 are chosen from 1 to 6 mol with 0.5 mol interval as well. In order to form NaBH4 solution with 7.5 wt.% NaBH4, 1 wt.% NaOH, 91.5 wt.% H2O, the molar quantities of NaBH4 are determined. For a PEM fuel cell operation, the required hydrogen production rates are estimated depending on 60, 65, 70 and 75 g of catalyst used in the NaBH4 solution at 25, 32.5 and 40 °C, respectively. It is concluded that the highest rate of hydrogen production per unit area from NaBH4 solution at 40 °C is found to be 3.834 × 10−5 g H2 s−1 cm−2 for 75 g catalyst. Utilizing 80% of this hydrogen production, the maximum amounts of power generation from a PEM fuel cell per unit area at 80 °C under 5 atm are estimated as 1.121 W cm−2 for 0.016 cm by utilizing hydrogen from 75 g catalyst assisted NaBH4 solution at 40 °C.  相似文献   

6.
《Journal of power sources》2006,158(2):1348-1357
Steam and autothermal reforming reactions of LPG (propane/butane) over high surface area CeO2 (CeO2 (HSA)) synthesized by a surfactant-assisted approach were studied under solid oxide fuel cell (SOFC) operating conditions. The catalyst provides significantly higher reforming reactivity and excellent resistance toward carbon deposition compared to the conventional Ni/Al2O3. These benefits of CeO2 are due to the redox property of this material. During the reforming process, the gas–solid reactions between the hydrocarbons present in the system (i.e. C4H10, C3H8, C2H6, C2H4, and CH4) and the lattice oxygen (OOx) take place on the ceria surface. The reactions of these adsorbed surface hydrocarbons with the lattice oxygen (CnHm + OOx  nCO + m/2(H2) + VO + 2e′) can produce synthesis gas (CO and H2) and also prevent the formation of carbon species from hydrocarbons decomposition reactions (CnHm  nC + 2mH2). Afterwards, the lattice oxygen (OOx) can be regenerated by reaction with the steam present in the system (H2O + VO + 2e′  OOx + H2). It should be noted that VO denotes as an oxygen vacancy with an effective charge 2+.At 900 °C, the main products from steam reforming over CeO2 (HSA) were H2, CO, CO2, and CH4 with a small amount of C2H4. The addition of oxygen in autothermal reforming was found to reduce the degree of carbon deposition and improve product selectivities by completely eliminating C2H4 formation. The major consideration in the autothermal reforming operation is the O2/LPG (O/C molar ratio) ratio, as the presence of a too high oxygen concentration could oxidize the hydrogen and carbon monoxide produced from the steam reforming. A suitable O/C molar ratio for autothermal reforming of CeO2 (HSA) was 0.6.  相似文献   

7.
《Journal of power sources》2005,145(2):712-715
We constructed a reformer of methane based on an electrochemical principle. This apparatus consists of the proton conducting ceramics electrolyte and the hydrogen-permeable metal membrane cathode. For methane reforming, a mixture of methane and oxygen gas is supplied to the porous Ag cathode. The hydrogen ions, which formed by the anode reaction: CH4 + O2  CO2 + 4H+ + 4e, are transported through the proton conducting ceramics to the cathode. Then, the hydrogen is formed at the cathode by the reaction: 4H+ + 4e  2H2. The hydrogen, which permeates through the metal membrane cathode, is 100% purity.The hydrogen separation ability of the reformer was investigated at 400–650 °C by measuring the electric current through the proton conducting oxide electrolyte. Since the ionic transport number of the proton conducting oxide is nearly unity, the current through the electrolyte corresponds to the proton flux through the electrolyte.The current measurements showed that the extracted proton flux through the electrolyte increased with increasing the applied voltage as well as temperature as we expected. However, the current measurements under the low voltage revealed that the extracted current was lesser than the expected value from Ohm's law. The decrease of the current is possibly caused for the reduction of the effective voltage by the anode polarization. In order to separate the hydrogen with higher efficiency, the applied voltage must be as low as possible using the thinner electrolyte and the improved anode.  相似文献   

8.
《Journal of power sources》2006,157(1):422-429
Reduced-temperature solid oxide fuel cells (SOFCs) – with thin Ce0.85Sm0.15O1.925 (SDC) electrolytes, thick Ni–SDC anode supports, and composite cathodes containing La0.6Sr0.4Co0.2Fe0.8O3 (LSCF) and SDC – were fabricated and tested with iso-octane/air fuel mixtures. An additional supported catalyst layer, placed between the fuel stream and the anode, was needed to obtain a stable output power density (e.g. 0.6 W cm−2 at 590 °C) without anode coking. The Ru-CeO2 catalyst produced CO2 and H2 at temperatures <350 °C, while H2 and CO became predominant above 500 °C. Power densities were substantially less than for the same cells with H2 fuel (e.g. 1.0 W cm−2 at 600 °C), due to the dilute (≈20%) hydrogen in the fuel mixture produced by iso-octane partial oxidation. Electrochemical impedance analysis showed a main arc that represented ≈60% of the total resistance, and that increased substantially upon switching from hydrogen to iso-octane/air.  相似文献   

9.
Ni–YSZ cermet anodes for solid oxide fuel cells (SOFCs) were fabricated at various sintering temperatures from NiO–YSZ composite particles made by spray pyrolysis (SP) technique. NiO particles covered with fine YSZ (Y2O3 stabilized ZrO2) particles were used as the composite particles, and the initial ratio of Ni and YSZ was set at 75:25 (mol%). As a result, the cermet anode sintered at 1350 °C showed the morphology in which fine YSZ grains were uniformly dispersed on the surface of Ni grain network. Electrical performance such as electrochemical activity and internal resistance of a Ni–YSZ cermet anode changed with sintering temperature. The anode fabricated at 1350 °C showed the highest electrical performance. Especially, a single cell voltage with the Ni–YSZ cermet anode kept very stable for 8000 h at 1000 °C in the SOFC operation condition of H2—3% H2O and air. The cermet anode after a long-term test had its initial morphology. It indicates that the Ni–YSZ cermet anode fabricated from NiO–YSZ composite particles is a very promising material for its practical use as SOFCs.  相似文献   

10.
《Journal of power sources》2006,155(2):353-357
Two types of solid oxide fuel cells (SOFCs), with thin Ce0.85Sm0.15O1.925 (SDC) or 8 mol% Y2O3-stabilized ZrO2 (YSZ) electrolytes, were fabricated and tested with iso-octane/air fuel mixtures. An additional Ru–CeO2 catalyst layer, placed between the fuel stream and the anode, was needed to obtain a stable output power density without anode coking. Thermodynamic analysis and catalysis experiments showed that H2 and CO were primary reaction products at ≈750 °C, but that these decreased and H2O and CO2 increased as the operating temperature dropped below ≈600 °C. Power densities for YSZ cells were 0.7 W cm−2 at 0.7 V and 790 °C, and for SDC cells were 0.6 W cm−2 at 0.6 V and 590 °C. Limiting current behavior was observed due to the relatively low (≈20%) H2 content in the reformed fuel.  相似文献   

11.
In this study, undoped ZrO2 thin films were deposited on single-crystal silicon substrates using liquid phase deposition. The undoped films were formed by hydrolysis of zirconium sulfate (Zr(SO4)2·4H2O) in the presence of H2O. A continuous oxide film was obtained by controlling adequate (NH4)2S2O8 concentration. The deposited films were characterized by SEM, FT-IR, XRD and DTA. Typically, the films showed excellent adhesion to the substrate with uniform particle diameter about 150 nm. The thicknesses of ZrO2 film were about 200 nm after 10 h deposition at 30 °C. These films shows single tetragonal phase after heat treated at 600 °C. High annealing temperature (e.g. 750 °C) may result in the phase transformation of (t)-ZrO2 into (m)-ZrO2.  相似文献   

12.
Applicability of gulfweed as feedstock for a biomass-to-liquid (BTL) process was studied for both production of gas with high syngas (CO + H2) content via gasification of gulfweed and removal of gaseous impurities using char obtained in the gasification. Gulfweed as aqueous biomass was gasified with He/CO2/O2 using a downdraft fixed-bed gasifier at ambient pressure and 900 °C at equivalence ratios (ER) of 0.1–0.3. The syngas content increased while the conversion to gas on a carbon basis decreased with decreasing ER. At an ER of 0.1 and He/CO2/O2 = 0/85/15%, the syngas content was maximized at 67.6% and conversion to gas on a carbon basis was 94.2%. The behavior of the desulfurization using char obtained during the gasification process at ER = 0.1 and He/CO2/O2 = 0/85/15% was investigated using a downdraft fixed-bed reactor at 250–550 °C under 3 atmospheres (H2S/N2, COS/N2, and a mixture of gases composed of CO, CO2, H2, N2, CH4, H2S, COS, and steam). The char had a higher COS removal capacity at 350 °C than commercial activated carbon because (Ca,Mg)S crystals were formed during desulfurization. The char simultaneously removed H2S and COS from the mixture of gases at 450 °C more efficiently than did activated carbon. These results support this novel BTL process consisting of gasification of gulfweed with CO2/O2 and dry gas cleaning using self-supplied bed material.  相似文献   

13.
HO2 and CH3 are major intermediate species presented during the oxidation of natural gas at intermediate temperatures and high pressures. Previous theoretical calculations have identified several product channels for HO2 and CH3 reactions, with CH3 + HO2  CH3O + OH and CH3 + HO2  CH4 + O2 being the dominant reaction pathways. Both reaction pathways play an important role in the kinetics of CH4 oxidation as CH3 + HO2  CH3O + OH is a chain-branching reaction whereas CH3 + HO2  CH4 + O2 a chain termination reaction.H2O2/CH4/Ar mixtures were shock-heated to a temperature between 1054 and 1249 K near 3.5 atm to initiate the reaction. OH radicals yielded from H2O2 thermal decomposition react with H2O2 and CH4 respectively to produce HO2 and CH3 in the reacting system. Using laser absorption spectroscopy, time-histories of H2O, OH and HO2 were measured behind reflected shock waves. The rate constant of reaction CH3 + HO2  CH3O + OH was determined to be 6.8 × 1012 cm3 mol?1 s?1 with an uncertainty factor of 1.4. The rate constant of the competing CH3 + HO2  CH4 + O2 reaction was determined to be 4.4 × 1012 cm3 mol?1 s?1, with an uncertainty factor of 2.1. In addition, the rate constants of two other major reactions of the reacting system, H2O2 (+M)  2OH (+M) and OH + CH4  CH3O + OH, were found to have excellent agreement with values recommended in literature.  相似文献   

14.
《Journal of power sources》2006,161(1):460-465
This paper describes results on direct-methane solid oxide fuel cell (air, LSM-YSZ|YSZ|Ni-YSZ, CH4) operation for combined electricity and syngas production. Thermodynamic equilibrium predictions showed that efficient methane conversion to syngas is expected for SOFC operating temperature >700 °C and O2−/CH4 ratios of ≈1. A simple thermal analysis was used to determine conditions where the cell produces enough heat to self-sustain its operating temperature; relatively low cell voltage and O2−/CH4 ratios > 1 were found to be useful. Fuel cells operated at T  750 °C, V  0.4 V, and O2−/CH4  1.2 yielded electrical power output of ∼0.7 W cm−2 and syngas production rates of ∼20 sccm cm−2. Stable cell operation without coking for >300 h was achieved.  相似文献   

15.
《Journal of power sources》2006,161(1):531-534
MgAl2O4 was synthesized through hydrolysis of metallic alkoxides of Mg2+ and Al3+. The formed spinel precursor phase was calcined at temperatures between 600 and 1100 °C, for 4 h. The spinel was utilized as a Ni/Pt catalyst support. The Ni/MgAl2O4 catalysts (15% Ni, w/w) containing small amounts of Pt were tested for methane steam reforming. The solids were analyzed by X-ray diffraction (XRD), temperature programmed reduction (TPR) with H2 and catalytic tests. The spinel phase was formed at temperatures above 700 °C. The addition of small amounts of Pt to Ni/MgAl2O4 promoted an increase in surface area. This probably caused the considerable increase in methane conversion.  相似文献   

16.
《Journal of power sources》2006,159(1):179-185
Spinel LiMn2O4 thin-film cathodes were obtained by spin-coating the chitosan-containing precursor solution on a Pt-coated silicon substrate followed by a two-stage heat-treatment procedure. The LiMn2O4 film calcined at 700 °C for 1 h showed the highest Li-ion diffusion coefficient, 1.55 × 10−12 cm2 s−1 (PSCA measurement) among all calcined films. It is attributed to the larger interstitial space and better crystal perfection of LiMn2O4 film calcined at 700 °C for 1 h. Consequently, the 700 °C-calcined LiMn2O4 film exhibited the best rate performance in comparison with the ones calcined at other temperatures.  相似文献   

17.
《Journal of power sources》2006,156(2):520-524
Because of the need for an efficient and inexpensive reforming catalyst, the objective of this work is to determine the feasibility of employing Mo2C catalyst for the steam reforming and oxy-steam reforming of the higher hydrocarbons typical of transportation fuels such as gasoline. It is shown that bulk Mo2C catalysts can successfully reform 2,2,4-trimethyl pentane (isooctane) to generate H2, CO and CO2 at very low steam/carbon ratios, without coke formation, eliminating the need for pre-reforming. Maximum hydrogen generation was observed at a S/C ratio of 1.3 and 1000 °C during SR reactions and S/C of 0.71, O2/C of 0.12 at 900 °C during oxidative steam reforming reactions.  相似文献   

18.
This study examined new innovative design of aluminum rectangular and triangular double-layered microchannel heat sink (RDLMCHS) and (TDLMCHS), respectively, using Al2O3–H2O and SiO2–H2O nanofluids. A series of experimental runs for different channel dimensions, different nanoparticles concentrations and types and several pumping powers showed excellent hydrothermal performance for DLMCHS over traditional single-layer (SLMCHS). The results showed that the sequential TDLMCHS provided a 27.4% reduction in the wall temperature comparing with RDLMCHS and has better temperature uniformity across the channel length with less than 2 °C. Sequential TDLMCHS provided 16.6% total thermal resistance lesser than the RDLMCHS at low pumping power and the given geometry parameters. Pressure drop observation showed no significant differences between the two designs. In addition, larger number of channels and smaller fin thickness referred less thermal resistance rather than only increasing the pumping power. Higher nanoparticle concentration showed better thermal stability for both nanofluids than pure water. The Al2O3–H2O nanofluid (0.9 vol.%) showed best performance with the temperature difference of 1.6 °C and lowest thermal resistance of 0.13 °C/W·m2.  相似文献   

19.
Activity and stability of FeTiO3, MnTiO3, NiFe2O4, CuFe2O4, NiCr2O4, 2CuO·Cr2O3, CuO and Fe2O3 for the atmospheric decomposition of concentrated sulfuric acid in sulfur-based thermochemical water splitting cycles are presented. Catalyst activity was determined at temperatures from 725 to 900 °C. Catalytic stability was examined at 850 °C for up to 1 week of continuous operation. The results were compared to a 1.0 wt% Pt/TiO2 catalyst. Surface area by nitrogen physisorption, X-ray diffraction analyses, and temperature programmed desorption and oxidation were used to characterize fresh and spent catalyst samples.Over the temperature range, the catalyst activity of the complex oxides followed the general trend: 2CuO·Cr2O3 > CuFe2O4 > NiCr2O4  NiFe2O4 > MnTiO3  FeTiO3. At temperatures less than 800 °C, the 1.0 wt% Pt/TiO2 catalyst had higher activity than the complex oxides, but at temperatures above 850 °C, the 2CuO·Cr2O3 and CuFe2O4 samples had the highest activity.Surface area was found to decrease for all of the metal oxides after exposure to reaction conditions. In addition, the two complex metal oxides that contained chromium were not stable in the reaction environment; both leached chromium into the acid stream and decomposed into their individual oxides. The FeTiO3 sample also produced a discoloration of the reactor due to minor leaching and converted to Fe2TiO5. Fe2O3, MnTiO3 and NiFe2O4 were relatively stable in the reaction environment. In addition, CuFe2O4 catalyst appeared relatively promising due to its high activity and lack of any leaching issues; however it deactivated in week-long stability experiments.Complex metal oxides may provide an attractive alternative to platinum-based catalyst for the decomposition of sulfuric acid; however, the materials examined in this study all displayed shortcomings including material sintering, phase changes, low activity at moderated temperatures due to sulfate formation, and decomposition to their individual oxides. More effort is needed in this area to discover metal oxide materials that are less expensive, more active and more stable than platinum catalysts.  相似文献   

20.
《Journal of power sources》2006,157(1):528-532
Single crystalline ammonium vanadium oxide bronze NH4V4O10 nanobelts were synthesized by the hydrothermal treatment of H2C2O4·2H2O and NH4VO3 at 140 °C for 48 h. The NH4V4O10 nanobelts were characterized using a combination of techniques including X-ray diffraction, transmission electron microscopy, selected area electronic diffraction, scanning electron microscopy and X-ray photoelectron spectroscopy techniques. The as-obtained nanobelts are several microns long, typically 30–40 nm wide, and 10–20 nm thick. The electrochemical properties of the nanobelts were tested in cells with metallic lithium as the negative electrode, the first discharge capacity of 171.8 mAh g−1 was achieved.  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