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 46 毫秒
1.
NaBH4 and KBH4 hydrolysis reactions (BH4 + 4H2O → B(OH)4 + 4H2), which can be utilized as a source of high purity hydrogen and be easily controlled catalytically, are exothermic processes. Precise determination of the evolved heat is of outmost importance for the design of the reactor for hydrogen generation. In this work we present an efficient calorimetric method for the direct measurement of the heats evolved during the catalyzed hydrolysis reaction. A modified Setaram Titrys microcalorimeter was used to determine the heat of hydrolysis in a system where water is added to pure solid NaBH4 or KBH4 as well as to solid NaBH4 or KBH4 mixed with a Co-based solid catalyst. The measured heats of NaBH4 hydrolysis reaction were: −236 kJ mol−1, −243 kJ mol−1, −235 kJ mol−1, and −236 kJ mol−1, without catalyst and in the presence of Co nanoparticles, CoO and Co3O4, respectively. In the case of the KBH4 hydrolysis reaction, the measured heats were: −220 kJ mol−1, −219 kJ mol−1, −230 kJ mol−1, and −228 kJ mol−1, without catalyst and with Co nanoparticles, CoO and Co3O4, respectively. Also, a comparison was made with an aqueous solution of CoCl2·6H2O used as catalyst in which case the measured heats were −222 kJ mol−1 and −196 kJ mol−1 for NaBH4 and KBH4 hydrolysis, respectively. The influence of solid NaOH or KOH additions on the heat of borohydride hydrolysis has been investigated as well.  相似文献   

2.
Chitosan (Chs) flakes were prepared from chitin materials that were extracted from the exoskeleton of Cape rock lobsters in South Africa. The Chs flakes were prepared into membranes and the Chs membranes were modified by cross-linking with H2SO4. The cross-linked Chs membranes were characterized for the application in direct methanol fuel cells. The Chs membrane characteristics such as water uptake, thermal stability, proton resistance and methanol permeability were compared to that of high performance conventional Nafion 117 membranes. Under the temperature range studied 20-60 °C, the membrane water uptake for Chs was found to be higher than that of Nafion. Thermal analysis revealed that Chs membranes could withstand temperature as high as 230 °C whereas Nafion 117 membranes were stable to 320 °C under nitrogen. Nafion 117 membranes were found to exhibit high proton resistance of 284 s cm−1 than Chs membranes of 204 s cm−1. The proton fluxes across the membranes were 2.73 mol cm−2 s−1 for Chs- and 1.12 mol cm−2 s−1 Nafion membranes. Methanol (MeOH) permeability through Chs membrane was less, 1.4 × 10−6 cm2 s−1 for Chs membranes and 3.9 × 10−6 cm2 s−1 for Nafion 117 membranes at 20 °C. Chs and Nafion membranes were fabricated into membrane electrode assemblies (MAE) and their performances measure in a free-breathing commercial single cell DMFC. The Nafion membranes showed a better performance as the power density determined for Nafion membranes of 0.0075 W cm−2 was 2.7 times higher than in the case of Chs MEA.  相似文献   

3.
Ammonia borane (AB) hydrolysis is a potential process for on-board hydrogen generation. This paper presents isothermal hydrogen release rate measurements of dilute AB (1 wt%) hydrolysis in the presence of carbon supported ruthenium catalyst (Ru/C). The ranges of investigated catalyst particle sizes and temperature were 20-181 μm and 26-56 °C, respectively. The obtained rate data included both kinetic and diffusion-controlled regimes, where the latter was evaluated using the catalyst effectiveness approach. A Langmuir-Hinshelwood kinetic model was adopted to interpret the data, with intrinsic kinetic and diffusion parameters determined by a nonlinear fitting algorithm. The AB hydrolysis was found to have an activation energy 60.4 kJ mol−1, pre-exponential factor 1.36 × 1010 mol (kg-cat)−1 s−1, adsorption energy −32.5 kJ mol−1, and effective mass diffusion coefficient 2 × 10−10 m2 s−1. These parameters, obtained under dilute AB conditions, were validated by comparing measurements with simulations of AB consumption rates during the hydrolysis of concentrated AB solutions (5-20 wt%), and also with the axial temperature distribution in a 0.5 kW continuous-flow packed-bed reactor.  相似文献   

4.
The temporal variation of OH (A2Σ+) chemiluminescence in hydrogen oxidation chemistry has been studied in a shock tube behind reflected shock waves at temperatures of 1400-3300 K and at a pressure of 1 bar. The aim of the present work is to obtain a validated reaction scheme to describe OH formation in the H2/O2 system. Temporal OH emission profiles and ignition delay times for lean and stoichiometric H2/O2 mixtures diluted in 97-98% argon were obtained from the shock-tube experiments. Based on a literature review for the hydrogen combustion system, the key reaction considered was H + O + M = OH + M (R1). The temperature dependence of the measured peak OH emission from the shock tube and the peak OH concentration from a homogeneous closed reactor model are compared. Based on these results a reaction rate coefficient of k1 = (1.5 ± 0.4) × 1013 exp(−25 kJ mol−1/RT) cm6 mol−2 s−1 was found for the forward reaction (R1) which is slightly higher than the rate coefficient suggested by Hidaka et al. (1982). The comparison of measured and simulated absolute concentrations shows good agreement. Additionally, a one-dimensional laminar premixed low-pressure flame calculation was performed for where absolute OH concentration measurements have been reported by Smith et al. (2005). The absolute peak OH concentration is fairly well reproduced if the above mentioned rate coefficient is used in the simulation.  相似文献   

5.
This work presents the kinetic and thermodynamic studies of the Bunsen reaction in the sulfur–iodine thermochemical cycle for hydrogen production by water splitting. A series of experimental runs have been carried out by feeding the gas mixture SO2/N2 in an I2/H2O medium in the temperature range of 336–358 K. The effects of the various operating parameters on the SO2 conversion ratio have been evaluated. The results showed that the efficiency of SO2 conversion into H2SO4 increased with the amount of I2 or H2O increase. The increasing reaction temperature impeded SO2 conversion into H2SO4. A kinetic model has been developed to fit to the experimental data obtained in a semi-batch reactor. A good fitting can be observed for each experiment, which discloses the overall kinetic mechanism of the complex Bunsen reaction. The apparent activation energies were found to be 23.513 kJ mol−1 and 9.212 kJ mol−1 for the sequential reactions  and , respectively.  相似文献   

6.
The thermodynamic properties of CeMn1−xAl1−xNi2x (x=0.00, 0.25, 0.50 and 0.75) hydrides have been investigated in this paper. With increasing Ni substitution content, the hydrogen concentration (H/M) in CeMn1−xAl1−xNi2x (x=0.00, 0.25, 0.50 and 0.75) hydride increases from 0.129 wt% for x=0.00 to 0.421 wt% for x=0.75 at 293 K. The pressure–concentration isotherm (P–C–T) curves show that no hydrogen equilibrium pressure plateau has been observed for CeMnAl hydride while the slope of the plateau become flatter and longer with increasing Ni content. Meanwhile, the enthalpy change (ΔH0) and the entropy change (ΔS0) of the hydrides for dehydrogenization shift from −67.44 kJ mol−1 (x=0.00) to 21.16 kJ mol−1 (x=0.75) and from −0.24 kJ mol−1 K−1 (x=0) to −0.03 kJ mol−1 K−1 (x=0.75), respectively. With increasing Ni content, both ΔH0 and ΔS0 for dehydrogenization shift to the positive direction and make alloy hydrides more stable and hydrogen desorption much easier.  相似文献   

7.
The kinetic parameters of carbon monoxide and methanol oxidation reactions on a high performance carbon-supported Pt-Ru electrocatalyst (HP 20% 1:1 Pt-Ru alloy on Vulcan XC-72 carbon black) have been studied using cyclic voltammetry and rotating disk electrode (RDE) techniques in 0.50 M H2SO4 and H2SO4 (0.06-0.92 M) + CH3OH (0.10-1.00 M) solutions at 25.0-45.0 °C. CO oxidation showed an irreversible behaviour with an adsorption control giving an exchange current density of 2.3 × 10−6 A cm−2 and a Tafel slope of 113 mV dec−1 (α = 0.52) at 25.0 °C. Methanol oxidation behaved as an irreversible mixed-controlled reaction, probably with generation of a soluble intermediate (such as HCHO or HCOOH), showing an exchange current density of 7.4 × 10−6 A cm−2 and a Tafel slope of 199 mV dec−1 (α = 0.30) at 25.0 °C. Reaction orders of 0.5 for methanol and −0.5 for proton were found, which are compatible with the consideration of the reaction between Pt-CO and Ru-OH species as the rate-determining step, being the initial methanol adsorption adjustable to a Temkin isotherm. The activation energy calculated through Arrhenius plots was 58 kJ mol−1, practically independent of the applied potential. Methanol oxidation on carbon-supported Pt-Ru electrocatalyst was improved by multiple potential cycles, indicating the generation of hydrous ruthenium oxide, RuOxHy, which enhances the process.  相似文献   

8.
In this study, p(AMPS) hydrogels are synthesized from 2-acrylamido-2-methyl-1-propansulfonic acid (AMPS) via a photo polymerization technique. The hydrogels are used as template for metal nanoparticles and magnetic ferrite nanoparticles, and also as a catalysis vessel in the generation of hydrogen from the hydrolysis of NaBH4. Approximately 5 nm Ru (0) and 20-30 nm magnetic ferrite particles are generated in situ inside this p(AMPS) hydrogel network and then used as a catalysis medium in hydrogen production by hydrolysis of sodium boron hydride in a basic medium. With an applied external magnetic field, the hydrogel reactor, containing Ru and ferrite magnetic particles, can be removed from the catalysis medium; providing on-demand generation of hydrogen. The effect of various parameters such as the initial concentration of NaBH4, the amount of catalyst and temperature on the hydrolysis reaction is evaluated. The activation energy for hydrogen production by Ru (0) nanoparticles is found to be 27.5 kJ mol−1; while the activation enthalpy is 30.4 kJ mol−1. The hydrogen generation rate in presence of 5 wt% NaOH and 50 mg p(AMPS)-Ru catalyst is 8.2 L H2 min−1 g Ru.  相似文献   

9.
Ammonia borane (AB) is a candidate material for on-board hydrogen storage, and hydrolysis is one of the potential processes by which the hydrogen may be released. This paper presents hydrogen generation measurements from the hydrolysis of dilute AB aqueous solutions catalyzed by ruthenium supported on carbon. Reaction kinetics necessary for the design of hydrolysis reactors were derived from the measurements. The hydrolysis had reaction orders greater than zero but less than unity in the temperature range from 16 °C to 55 °C. A Langmuir–Hinshelwood kinetic model was adopted to interpret the data with parameters determined by a non-linear conjugate-gradient minimization algorithm. The ruthenium-catalyzed AB hydrolysis was found to have activation energy of 76 ± 0.1 kJ mol−1 and adsorption energy of −42.3 ± 0.33 kJ mol−1. The observed hydrogen release rates were 843 ml H2 min−1 (g catalyst)−1 and 8327 ml H2 min−1 (g catalyst)−1 at 25 °C and 55 °C, respectively. The hydrogen release from AB catalyzed by ruthenium supported on carbon is significantly faster than that catalyzed by cobalt supported on alumina. Finally, the kinetic rate of hydrogen release by AB hydrolysis is much faster than that of hydrogen release by base-stabilized sodium borohydride hydrolysis.  相似文献   

10.
In the present work, hydrogen generation through hydrolysis of a NaBH4(s)/catalyst(s) solid mixture was realized for the first time as a solid/liquid compact hydrogen storage system using Co nanoparticles as a model catalyst. The performance of the system was analysed from both the thermodynamic and kinetic points of view and compared with the classical catalyzed hydrolysis of a NaBH4 solution. The kinetic analysis of the NaBH4(s)/catalyst(s)/H2O(l) system shows that the reaction is first order with respect to the catalyst concentration, and the activation energy equal to 35 kJ molNaBH4−1. Additionally, calorimetric measurements of the heat evolved during the hydrolysis of NaBH4 solutions evidence the global process energy (−217 kJ molNaBH4−1). Characterization of the cobalt nanoparticles before and after the hydrolysis associated with the calorimetric measurements suggests the “in situ” formation of a catalytically active CoxB phase through “reduction” of an outer protective oxide layer that is regenerated at the end of reaction.  相似文献   

11.
In hydrogen solid–gas reaction at 300 K and 1 bar, the hydrogen content for Ti3.87Ni1.73Fe0.7Ox (0.2≤ × ≤0.8) alloys was in range 1.93–0.05 (Cwt.H,%), and discharge capacity of 360–235 A h/kg was achieved accordingly. The ΔHH2ΔHH2 and ΔSH2ΔSH2 values of −32.29 kJ mol−1 and −111.04 J mol−1 K−1, respectively, for Ti3.87Ni1.73Fe0.7O0.5 alloy were obtained using experimental PCT relations, where hysteresis effect was only slightly visible. The half-cell potentials (vs. Hg/HgO) of metal hydride (MH) electrodes based on Ti3.87Ni1.73Fe0.7Ox (0.2≤ ×≤ 0.8) alloys were calculated.  相似文献   

12.
The influence of multiple additions of two oxides, Cr2O3 and Nb2O5, as additives on the hydrogen sorption kinetics of MgH2 after milling was investigated. We found that the desorption kinetics of MgH2 were improved more by multiple oxide addition than by single addition. Even for the milled MgH2 micrometric size powders, the high hydrogen capacity with fast kinetics were achieved for the powders after addition of 0.2 mol% Cr2O3 + 1 mol% Nb2O5. For this composition, the hydride desorbed about 5 wt.% hydrogen within 20 min and absorbed about 6 wt.% in 5 min at 300 °C. Furthermore, the desorption temperature was decreased by 100 °C, compared to MgH2 without any oxide addition, and the activation energy for the hydrogen desorption was estimated to be about 185 kJ mol−1, while that for MgH2 without oxide was about 206 kJ mol−1.  相似文献   

13.
Interfacial lithium-ion transfer at the LiMn2O4 thin film electrode/aqueous solution was investigated. The cyclic voltammograms of the film electrode conducted in the aqueous solution was similar to an adsorption-type voltammogram of reversible system, suggesting that fast charge transfer reaction proceed in the aqueous solution system. We found that the activation energy for this interfacial lithium-ion transfer reaction obtains 23–25 kJ mol−1, which is much smaller than that in the propylene carbonate solution (50 kJ mol−1). This small activation energy will be responsible for the fast interfacial lithium-ion transfer reaction in the aqueous solution. These results suggest that fast lithium insertion/extraction reaction can be realized by decreasing the activation energy for interfacial lithium-ion transfer reaction.  相似文献   

14.
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   

15.
The mechanism of the oxygen reduction reaction (ORR) on nanoparticulated Pt/C-Nafion electrodes prepared in one step has been studied to simulate the reaction in the cathode of a Polymer Electrolyte Fuel Cell (PEFC). The kinetic parameters have been obtained by hydrodynamic polarization in O2-saturated 0.01–1.00 M H2SO4 and temperatures in the range 25.0–50.0 °C. The ORR current density was maximum and practically independent of the ionomer fraction in the rage 10–55 wt% Nafion. The poorer proton conductivity for lower Nafion fractions and the formation of catalyst areas completely surrounded by Nafion together with adsorption of Pt sites by sulfonate groups for higher Nafion fractions, explain the minor ORR activity in these conditions. The ionomer influence on the O2 diffusion at high overpotentials for Pt/C-Nafion was negligible when the Nafion content was smaller than 20 wt%. The higher kinetic current density for Pt/C-Nafion (100 mA cm−2) with respect to smooth Pt-Nafion (40 mA cm−2), together with the smaller activation energy of the former (25 ± 4 kJ mol−1) with respect to the latter (42 ± 5 kJ mol−1) highlighted the better properties attained by the nanosize effect. A remarkable novel result is that the reaction order of H+ in HClO4 is close to unity, whereas in sulfuric acid it is significantly smaller and changes with potential, what has been related to the sulfate adsorption. The anomalous dependence of the charge transfer coefficient with temperature was then explained by the thermal change of the double layer structure and the variation of the coverage of adsorbed species on Pt. The more sensitive effect for Pt/C-Nafion than for smooth Pt-Nafion was ascribed to the stronger interaction between the components when the nanoparticles are involved.  相似文献   

16.
We demonstrate a monolithic polymer electrolyte membrane fuel cell by integrating a narrow (200 μm) Nafion strip in a molded polydimethylsiloxane (PDMS) structure. We propose two designs, based on two 200 μm-wide and two 80 μm-wide parallel microfluidic channels, sandwiching the Nafion strip, respectively. Clamping the PDMS/Nafion assembly with a glass chip that has catalyst-covered Au electrodes, results in a leak-tight fuel cell with stable electrical output. Using 1 M CH3OH in 0.5 M H2SO4 solution as fuel in the anodic channel, we compare the performance of (I) O2-saturated 0.5 M H2SO4 and (II) 0.01 M H2O2 in 0.5 M H2SO4 oxidant solutions in the cathodic channel. For the 200 μm channel width, the fuel cell has a maximum power density of 0.5 mW cm−2 and 1.5 mW cm−2 at room temperature, for oxidants I and II, respectively, with fuel and oxidant flow rates in the 50-160 μL min−1 range. A maximum power density of 3.0 mW cm−2 is obtained, using oxidant II for the chip with 80 μm-wide channel, due to an improved design that reduces oxidant and fuel depletion effects near the electrodes.  相似文献   

17.
Heterogeneous (TiO2/UV, TiO2/H2O2/UV) and homogenous (H2O2/UV, Fe2+/H2O2/UV) solar advanced oxidation processes (AOPs) are proposed for the treatment of recalcitrant textile wastewater at pilot-plant scale with compound parabolic collectors (CPCs). The textile wastewater presents a lilac colour, with a maximum absorbance peak at 516 nm, high pH (pH = 11), moderate organic content (DOC = 382 mg C L−1, COD = 1020 mg O2 L−1) and high conductivity (13.6 mS cm−1), associated with a high concentration of chloride (4.7 g Cl L−1). The DOC abatement is similar for the H2O2/UV and TiO2/UV processes, corresponding only to 30% and 36% mineralization after 190 kJUV L−1. The addition of H2O2 to TiO2/UV system increased the initial degradation rate more than seven times, leading to 90% mineralization after exposure to 100 kJUV L−1. All the processes using H2O2 contributed to an effective decolourisation, but the most efficient process for decolourisation and mineralization was the solar-photo-Fenton with an optimum catalyst concentration of 100 mg Fe2+ L−1, leading to 98% decolourisation and 89% mineralization after 7.2 and 49.1 kJUV L−1, respectively. According to the Zahn-Wellens test, the energy dose necessary to achieve a biodegradable effluent after the solar-photo-Fenton process with 100 mg Fe2+ L−1 is 12 kJUV L−1.  相似文献   

18.
Magnesium-based alloys are among the promising materials for hydrogen storage and fuel cell applications due to their high hydrogen content. In the present work, we investigated the hydrogen release/uptake properties of the Mg–Ti–H system. Samples were prepared from the mixtures of MgH2 and TiH2 in molar ratios of 7:1 and 4:1 using a high-energy-high-pressure (HEHP) mechanical ball-milling method under 13.8 MPa hydrogen pressure. Thermogravimetric analysis (TGA) showed that a relatively large amount of hydrogen (5.91 and 4.82 wt.%, respectively, for the above two samples) was released between 126 and 313 °C while temperature was increased at a heating rate of 5 °C min−1 under an argon flow. The onset dehydrogenation temperature of these mixtures, which is 126 °C, is much lower than that of MgH2 alone, which is 381 °C. The activation energy of dehydrogenation was 71 kJ mol−1, which is much smaller than that of as-received MgH2 (153 kJ mol−1) or as-milled MgH2 (96 kJ mol−1). Furthermore, the hydrogen capacity and the dehydrogenation temperature remained largely unchanged over five dehydrogenation and rehydrogenation cycles.  相似文献   

19.
The aqueous-phase reforming (APR) of n-butanol (n-BuOH) over Ni(20 wt%) loaded Al2O3 and CeO2 catalysts has been studied in this paper. Over 100 h of run time, the Ni/Al2O3 catalyst showed significant deactivation compared to the Ni/CeO2 catalyst, both in terms of production rates and the selectivity to H2 and CO2. The Ni/CeO2 catalyst demonstrated higher selectivity for H2 and CO2, lower selectivity to alkanes, and a lower amount of C in the liquid phase compared to the Ni/Al2O3 sample. For the Ni/Al2O3 catalyst, the selectivity to CO increased with temperature, while the Ni/CeO2 catalyst produced no CO. For the Ni/CeO2 catalyst, the activation energies for H2 and CO2 production were 146 and 169 kJ mol−1, while for the Ni/Al2O3 catalyst these activation energies were 158 and 175 kJ mol−1, respectively. The difference of the active metal dispersion on Al2O3 and CeO2 supports, as measured from H2-pulse chemisorption was not significant. This indicates deposition of carbon on the catalyst as a likely cause of lower activity of the Ni/Al2O3 catalyst. It is unlikely that carbon would build up on the Ni/CeO2 catalyst due to higher oxygen mobility in the Ni doped non-stoichiometric CeO2 lattice. Based on the products formed, the proposed primary reaction pathway is the dehydrogenation of n-BuOH to butaldehyde followed by decarbonylation to propane. The propane then partially breaks down to hydrogen and carbon monoxide through steam reforming, while CO converts to CO2 mostly through water gas shift. Ethane and methane are formed via Fischer-Tropsch reactions of CO/CO2 with H2.  相似文献   

20.
A new process for generating hydrogen via near room temperature hydrolysis of AB complex using small amounts of platinum group metal catalyst has been studied. Using in situ 11B NMR spectroscopy, the overall rate of K2Cl6Pt catalyzed hydrolysis of AB complex was calculated to be third-order. The pre-exponential factor (A) and the activation energy (Ea) of Arrhenius equation, ln k = ln A − Ea/RT, were determined to be: A = 1.6 × 1011 L mol−1 s−1 and Ea = 86.6 kJ mol−1 for temperature range of (25–35 °C). X-ray photoelectron spectroscopy of the residue suggested that the platinum salt was reduced from Pt4+ to Pt0 within the course of the reaction and X-ray diffraction analysis pattern for the residue showed crystallized single-phase boric acid.  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