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.
The purpose of this study was to investigate the production of NaBH4 as hydrogen storage material and testing its hydrogen storage capacity in presence of catalyst. Synthesis of NaBH4 was investigated with NaH and trimethyl borate which was also produced in previous studies. Different reaction temperatures, times and reactant ratios constitute the three parameters of the production process. The best combination determined by FT-IR, XRF and XRD analyses, was found to be 1.413 (mol/mol) TMB over NaH at 250 °C for 90 min. In order to increase the NaBH4 purity ethylene diamine was used as solvent at 75 °C. After 4 extraction and crystallization processes 85.17% NaBH4 purity was obtained. Moreover, hydrogen content of the product NaBH4 was measured in a system with different catalysts since catalyst efficiency is important in decreasing the water dependence of dehydrogenation. CoCl2 catalyst proved best by taking out more hydrogen, both from the NaBH4 structure and from water.  相似文献   

2.
MgH2 is one of the most promising hydrogen storage materials due to its high capacity and low cost. In an effort to develop MgH2 with a low dehydriding temperature and fast sorption kinetics, doping MgH2 with NiCl2 and CoCl2 has been investigated in this paper. Both the dehydrogenation temperature and the absorption/desorption kinetics have been improved by adding either NiCl2 or CoCl2, and a significant enhancement was obtained in the case of the NiCl2 doped sample. For example, a hydrogen absorption capacity of 5.17 wt% was reached at 300 °C in 60 s for the MgH2/NiCl2 sample. In contrast, the ball-milled MgH2 just absorbed 3.51 wt% hydrogen at 300 °C in 400 s. An activation energy of 102.6 kJ/mol for the MgH2/NiCl2 sample has been obtained from the desorption data, 18.7 kJ/mol and 55.9 kJ/mol smaller than those of the MgH2/CoCl2, which also exhibits an enhanced kinetics, and of the pure MgH2 sample, respectively. In addition, the enhanced kinetics was observed to persist even after 9 cycles in the case of the NiCl2 doped MgH2 sample. Further kinetic investigation indicated that the hydrogen desorption from the milled MgH2 is controlled by a slow, random nucleation and growth process, which is transformed into two-dimensional growth after NiCl2 or CoCl2 doping, suggesting that the additives reduced the barrier and lowered the driving forces for nucleation.  相似文献   

3.
The study focuses on hydrogen production from steam reforming of glycerol over nickel based catalyst promoted by zirconia and supported over ceria. Catalyst was prepared by the wet-impregnation method and characterized by BET surface area analysis, X-ray diffraction technique and scanning electron microscopy (SEM) analysis. The performance of the catalyst was evaluated in terms of hydrogen yield, selectivity and glycerol conversion at 700 °C in a tubular fixed bed reactor. The effect of glycerol concentration in feed, space time (W/FAO), temperature and time on stream (TOS) was analyzed for the catalyst Ni–ZrO2/CeO2 which showed the complete conversion of glycerol and high H2 yield that corresponds to 3.95 mol of H2 out of 7 mol. Thermodynamic analysis was also carried out using Aspen HYSYS for system having glycerol concentration 10 wt% and 20 wt% and experimental results were compared with thermodynamics. Kinetic study was carried out for the steam reforming of glycerol over Ni–ZrO2/CeO2 catalyst using the power law model. The values of activation energy and order of reaction were found to be 43.4 kJ/mol and 0.3 respectively.  相似文献   

4.
The esterification of oleic acid in subcritical methanol catalyzed by zinc acetate was investigated in a batch-type autoclave. The effect of reaction conditions such as temperature, pressure, reaction time and molar ratio of oleic acid to methanol on the esterification was examined. The oleic acid conversion reached 95.0% under 220 °C and 6.0 MPa with the molar ratio of methanol to oleic acid being 4 and 1.0 wt% zinc acetate as catalyst. A kinetic model for the esterification was established. By fitting the kinetic model with the experimental results, the reaction order n = 2.2 and activation energy Ea = 32.62 KJ/mol were obtained.  相似文献   

5.
The kinetic parameters for the “inversion substitution” reactions of Br-containing fire suppressants CH3Br and CF3Br with molecular oxygen CX3Br + O2  CX3O2 + Br (X = H, F) have been obtained on the basis of computational study carried out by quantum chemistry methods and transition state theory (TST). The DFT/B3LYP approach has been used to obtain molecular structures and vibrational frequencies and CCSD(T) method with additive corrections at MP2 and SODFT levels have been applied in order to calculate reaction enthalpies and activation barriers. Activation barrier for the substitution reaction CH3Br + O2  CH3O2 + Br was found to be (53.3 kcal/mol) close to the endoergicity (51.2 kcal/mol) of bimolecular abstraction CH3Br + O2  CH2Br + HO2 which means that this “inversion substitution” should be considered as a competing process in the oxidation initiation of bromine containing substances. The results of these calculations together with the earlier results have allowed us to build the correlation of Evans–Polanyi type between the activation barrier and reaction enthalpy for “inversion substitution” reactions CH3X + O2  CH3O2 + X. For bromotrifluoromethane CF3Br (Halon-1301) the barrier of “inversion substitution” was found to be of 86 kcal/mol that is higher by about 15 kcal/mol as compared with unimolecular decay CF3Br  CF3 + Br barrier. This makes negligible the role of bimolecular substitution reaction for kinetics of CF3Br in oxygen environment. The TST calculations of the rate constants and their temperature dependence for the direct and reverse “inversion substitution” reactions have been carried out within the temperature interval of 273–2000 K.  相似文献   

6.
Al2O3–ZrO2 (AZ) xerogel supports prepared by a sol-gel method were calcined at various temperatures. Ni/Al2O3–ZrO2 (Ni/AZ) catalysts were then prepared by an impregnation method for use in hydrogen production by steam reforming of liquefied natural gas (LNG). The effect of calcination temperature of AZ supports on the catalytic performance of Ni/AZ catalysts in the steam reforming of LNG was investigated. Crystalline phase of AZ supports was transformed in the sequence of amorphous γ-Al2O3 and amorphous ZrO2  θ-Al2O3 and tetragonal ZrO2   + α)-Al2O3 and (tetragonal + monoclinic) ZrO2  α-Al2O3 and (tetragonal + monoclinic) ZrO2 with increasing calcination temperature from 700 to 1300 °C. Nickel oxide species were strongly bound to γ-Al2O3 and θ-Al2O3 in the Ni/AZ catalysts through the formation of solid solution. In the steam reforming of LNG, both LNG conversion and hydrogen composition in dry gas showed volcano-shaped curves with respect to calcination temperature of AZ supports. Nickel surface area of Ni/AZ catalysts was well correlated with catalytic performance of the catalysts. Among the catalysts tested, Ni/AZ1000 (nickel catalyst supported on AZ support that had been calcined at 1000 °C) with the highest nickel surface area showed the best catalytic performance. Well-developed and pure tetragonal phase of ZrO2 in the AZ1000 support played an important role in the adsorption of steam and the subsequent spillover of steam from the support to the active nickel.  相似文献   

7.
Experimental measurements of burning rates, analysis of the key reactions and kinetic pathways, and modeling studies were performed for H2/CO/O2/diluent flames spanning a wide range of conditions: equivalence ratios from 0.85 to 2.5, flame temperatures from 1500 to 1800 K, pressures from 1 to 25 atm, CO fuel fractions from 0 to 0.9, and dilution concentrations of He up to 0.8, Ar up to 0.6, and CO2 up to 0.4. The experimental data show negative pressure dependence of burning rate at high pressure, low flame temperature conditions for all equivalence ratios and CO fractions as high as 0.5. Dilution with CO2 was observed to strengthen the pressure and temperature dependence compared to Ar-diluted flames of the same flame temperature. Simulations were performed to extend the experimentally studied conditions to conditions typical of gas turbine combustion in Integrated Gasification Combined Cycle processes, including preheated mixtures and other diluents such as N2 and H2O.Substantial differences are observed between literature model predictions and the experimental data as well as among model predictions themselves – up to a factor of three at high pressures. The present findings suggest the need for several rate constant modifications of reactions in the current hydrogen models and raise questions about the sufficiency of the set of hydrogen reactions in most recent hydrogen models to predict high pressure flame conditions relevant to controlling NOx emissions in gas turbine combustion. For example, the reaction O + OH + M = HO2 + M is not included in most hydrogen models but is demonstrated here to significantly impact predictions of lean high pressure flames using rates within its uncertainty limits. Further studies are required to reduce uncertainties in third body collision efficiencies for and fall-off behavior of H + O2(+M) = HO2(+M) in both pure and mixed bath gases, in rate constants for HO2 reactions with other radical species at higher temperatures, and in rate constants for reactions such as O + OH + M that become important under the present conditions in order to properly characterize the kinetics and predict global behavior of high-pressure H2 or H2/CO flames.  相似文献   

8.
Numerical results of the Navier–Stokes equations and the DSMC method for H2/N2 and H2/N2/CO2 mixtures in planar microchannels are compared in order to verify the accuracy of slip/jump boundary conditions in multicomponent flows. For the solution of the Navier–Stokes and energy equations, a finite-volume method is used together with a complete set of slip/jump boundary conditions derived from the kinetic theory of gases. In order to compare the accuracy of these two methods, different wall temperatures as well as inlet mass flow rates are examined. Isothermal flow is also considered for comparison with available analytical solutions for slip flows. The Knudsen number typically ranged between 0.015 < Kn < 0.09 with a Reynolds number of 5 < Re < 15. The two methods are generally in good agreement under the conditions studied especially at low to moderate Knudsen numbers.  相似文献   

9.
The energy and exergy analyses of the drying process of olive mill wastewater (OMW) using an indirect type natural convection solar dryer are presented. Olive mill wastewater gets sufficiently dried at temperatures between 34 °C and 52 °C. During the experimental process, air relative humidity did not exceed 58%, and solar radiation ranged from 227 W/m2 to 825 W/m2. Drying air mass flow was maintained within the interval 0.036–0.042 kg/s. Under these experimental conditions, 2 days were needed to reduce the moisture content to approximately one-third of the original value, in particular from 3.153 gwater/gdry matter down to 1.000 gwater/gdry matter.Using the first law of thermodynamics, energy analysis was carried out to estimate the amounts of energy gained from solar air heater and the ratio of energy utilization of the drying chamber. Also, applying the second law, exergy analysis was developed to determine the type and magnitude of exergy losses during the solar drying process. It was found that exergy losses took place mainly during the second day, when the available energy was less used. The exergy losses varied from 0 kJ/kg to 0.125 kJ/kg for the first day, and between 0 kJ/kg and 0.168 kJ/kg for the second. The exergetic efficiencies of the drying chamber decreased as inlet temperature was increased, provided that exergy losses became more significant. In particular, they ranged from 53.24% to 100% during the first day, and from 34.40% to 100% during the second.  相似文献   

10.
The pyrolysis of cyclohexane at low pressure (40 mbar) was studied in a plug flow reactor from 950 to 1520 K by synchrotron VUV photoionization mass spectrometry. More than 30 species were identified by measurement of photoionization efficiency (PIE) spectra, including some radicals like methyl, propargyl, allyl and cyclopentadienyl radicals, and stable products (e.g., 1-hexene, benzene and some aromatics). Among all the products, 1-hexene is formed at the lowest temperature, indicating that the isomerization of cyclohexane to 1-hexene is the dominant initial decomposition channel under the condition of our experiment. We built a kinetic model including 148 species and 557 reactions to simulate the experimental results. The model satisfactorily reproduced the mole fraction profiles of most pyrolysis products. The rate of production (ROP) analysis at 1360 and 1520 K shows that cyclohexane is consumed mainly through two reaction sequences: cyclohexane  1-hexene  allyl radical + n-propyl radical, and cyclohexane  cyclohexyl radical  hex-5-en-1-yl radical that further decomposes to 1,3-butadiene via hex-1-en-3-yl and but-3-en-1-yl radicals. Besides the stepwise dehydrogenation of cyclohexane, C3 + C3 channels, i.e. C3H3 + C3H3 and C3H3 + aC3H5 also have important contribution to benzene formation. The simulation reveals that C3H3 + C3H3 = phenyl + H reaction is the key step for other aromatics formation, i.e. toluene, phenylacetylene, styrene, ethylbenzene and indene in this work.  相似文献   

11.
The reaction pathways for the oxidation by O2 of polycyclic aromatic hydrocarbons present in soot particles are investigated using density functional theory at B3LYP/6-311++G(d,p) level of theory. For this, pyrene radical (4-pyrenyl) is chosen as the model molecule, as most soot models present in the literature employ the reactions involving the conversion of 4-pyrenyl to 4-phenanthryl by O2 and OH to account for soot oxidation. Several routes for the formation of CO and CO2 are proposed. The addition of O2 on a radical site to form a peroxyl radical is found to be barrierless and exothermic with reaction energy of 188 kJ/mol. For the oxidation reaction to proceed further, three pathways are suggested, each of which involve the activation energies of 104, 167 and 115 kJ/mol relative to the peroxyl radical. The effect of the presence of H atom on a carbon atom neighboring the radical site on the energetics of carbon oxidation is assessed. Those intermediate species formed during oxidation with seven-membered rings or with a phenolic group are found to be highly stable. The rate constants evaluated using transition state theory in the temperature range of 300–3000 K for the reactions involved in the mechanism are provided.  相似文献   

12.
《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   

13.
The influence of organic loading rates (OLRs) on the performance of fermentative hydrogen-producing bioreactors operating in continuous stirred tank reactor (CSTR) and membrane bioreactor (MBR) modes was examined. Five OLRs were examined, ranging from 4.0 to 30 g COD L?1 d?1, with influent glucose concentrations ranging from 1.3 to 10 g COD L?1. At OLRs up to 13 g COD L?1 d?1, all influent glucose was utilized and the H2 yield was not significantly influenced by OLR, although the yield in the CSTR mode was significantly higher than that in the MBR mode, 1.25 versus 0.97 mol H2 (mol Gluc. Conv.)?1, respectively. At an OLR of 30 g COD L?1 d?1, both reactor modes were overloaded with respect to glucose utilization and also had significantly higher H2 yields of 1.77 and 1.49 mol H2 (mol Gluc. Conv.)?1 for the CSTR and MBR modes, respectively, versus the underloaded operation. At the intermediate OLR of 22 g COD L?1 d?1, the H2 yield was maximized at 1.78 mol H2 (mol Gluc. Conv.)?1 for both the CSTR and MBR operation. Overall H2 production was 50% higher in the MBR mode, 0.78 versus 0.51 moles d?1, because the CSTR mode was overloaded with respect to glucose utilization at this OLR. These results suggest that an optimum OLR that maximizes H2 yield and H2 production may be near the OLR that causes overload with respect to substrate utilization. Additionally, while the CSTR mode is easier to operate and provides higher H2 yields at underloaded and overloaded OLRs, the MBR mode may be preferable when operating near the optimum OLR.  相似文献   

14.
《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   

15.
A conceptual design is presented for a hybrid sulfur process for the production of hydrogen using a high-temperature nuclear heat source to split water. The process combines proton exchange membrane-based SO2-depolarized electrolyzer technology being developed at Savannah River National Laboratory with silicon carbide bayonet decomposition reactor technology being developed at Sandia National Laboratories. Both are part of the US DOE Nuclear Hydrogen Initiative. The flowsheet otherwise uses only proven chemical process components. Electrolyzer product is concentrated from 50 wt% sulfuric acid to 75 wt% via recuperative vacuum distillation. Pinch analysis is used to predict the high-temperature heat requirement for sulfuric acid decomposition. An Aspen Plus? model of the flowsheet indicates 340.3 kJ high-temperature heat, 75.5 kJ low-temperature heat, 1.31 kJ low-pressure steam, and 120.9 kJ electric power are consumed per mole of H2 product, giving an LHV efficiency of 35.3% (41.7% HHV efficiency) if electric power is available at a conversion efficiency of 45%.  相似文献   

16.
The aim of this work was an experimental study of the temporal evolution of isobaric adsorption uptake (release) for simplest configuration of an adsorbent-heat exchanger unit, namely, a monolayer of loose adsorbent grains located on a metal plate. The study was performed by a large temperature jump method at four various boundary conditions of an adsorptive heat transformation cycle typical for air-conditioning application driven by low temperature heat: Te = 5 and 10 °C, Tc = 30 and 35 °C and THS = 80 °C. The size of the Fuji silica grains was varied from 0.2 to 1.8 mm to investigate its effect on water sorption dynamics. For each boundary set and grain size the experimental kinetic curve could be described by an exponential function up to 80–90% of the equilibrium conversion. Desorption runs are found to be faster than appropriate adsorption runs by a factor of 2.2–3.5, hence, for optimal durations of the isobaric ad- and desorption phases of the chilling cycle should be selected accordingly. The size R of the adsorbent grains was found to be a powerful tool to manage the dynamics of isobaric water ad-/desorption. For large grains the characteristic time was strongly dependent on the grain size and proportional to R2. Much less important appeared to be an impact of the boundary conditions which variation just weakly affected the dimensionless kinetic curves for the four tested cycles. The maximal specific cooling/heating power was proportional to the maximal temperature difference ΔT and the contact area S between the layer and the metal plate, and can exceed 10 kW/kg. The heat transfer coefficient α estimated from this power was as large as 100–250 W/(m2 K) that much exceeds the value commonly used to describe the cycle dynamics.  相似文献   

17.
Mg2In0.1Ni solid solution with an Mg2Ni-type structure has been synthesized and its hydrogen storage properties have been investigated. The results showed that the introduction of In into Mg2Ni not only significantly improved the dehydrogenation kinetics but also greatly lowered the thermodynamic stability. The dehydrogenation activation energy (Ea) and enthalpy change (ΔH) decreased from 80 kJ/mol and 64.5 kJ/mol H2 to 28.9 kJ/mol and 38.4 kJ/mol H2, respectively. The obtained results point to a method for improving not only the thermodynamic but also the kinetic properties of hydrogen storage materials.  相似文献   

18.
This paper investigates the effects of various fuels on hydrogen production for automotive PEM fuel cell systems. Gasoline, methanol, ethanol, dimethyl ether and methane are compared for their effects on fuel processor size, start-up energy and overall efficiencies for 50 kWe fuel processors. The start-up energy is the energy required to raise the temperature of the fuel processor from ambient temperature (20 °C) to that of the steady-state operating temperatures. The fuel processor modeled consisted of an equilibrium-ATR (autothermal), high-temperature water gas shift (HTS), low-temperature water gas shift (LTS) and preferential oxidation (PrOx) reactors. The individual reactor volumes with methane, dimethyl ether, methanol and ethanol were scaled relative to a gasoline-fueled fuel processor meeting the 2010 DOE technical targets. The modeled fuel processor volumes were, 25.9 L for methane, 30.8 L for dimethyl ether, 42.5 L for gasoline, 43.7 L for ethanol and 45.8 L for methane. The calculated fuel processor start-up energies for the modeled fuels were, 2712 kJ for methanol, 3423 kJ for dimethyl ether, 6632 kJ for ethanol, 7068 kJ for gasoline and 7592 kJ for methane. The modeled overall efficiencies, correcting for the fuel processor start-up energy using a drive cycle of 33 miles driven per day, were, 38.5% for dimethyl ether, 38.3% for methanol, 37% for gasoline, 34.5% for ethanol and 33.2% for methane assuming a steady-state efficiency of 44% for each fuel.  相似文献   

19.
The flame structure and kinetics of dimethyl ether (DME) flames with and without CO2 dilution at reduced and elevated pressures were studied experimentally and computationally. The species distributions of DME oxidation in low-pressure premixed flat flames were measured by using electron-ionization molecular-beam mass spectrometry (EI-MBMS) at an equivalence ratio of 1.63 and 50 mbar. High-pressure flame speeds of lean and rich DME flames with and without CO2 dilution were measured in a nearly-constant-pressure vessel between about 1 and 20 bar. The experimental results were compared with predictions from four kinetic models: the first was published by Zhao et al. (2008) [9], the second developed by the Lawrence Livermore National Laboratory (LLNL) (Kaiser et al., 2000) [13], and the third has been made available to us as the Aramco mechanism (Metcalfe et al., 2013) [14]; as the fourth, we have used an updated model developed in this study. Good agreement was found between measurements and predictions from all four models for all major and most typical intermediate species with and without CO2 addition in low-pressure flat flame experiments. However, none of the models was able to reliably predict high-pressure flame speeds. Although the updated model improved the prediction of flame speeds for lean mixtures, errors remained for rich conditions at elevated pressure, likely due to uncertainty in the rates of CH3 + H(+M) = CH4(+M) and the branching and termination reaction pair of CH3 + HO2 = CH3O + OH and CH3 + HO2 = CH4 + O2. CO2 addition considerably decreased the flame speed. Kinetic comparisons between inert and chemically active CO2 in DME flames showed that CO2 addition affects rich and lean DME flame kinetics differently. For lean flames, both the inert third-body effect and the kinetic effect of CO2 reduce H-atom production. However, for rich flames, the inert third-body effect increases H-atom production via HCO(+M) = H + CO(+M) and suppression of the kinetic effect of CO2 by shifting the equilibrium of CO + OH = CO2 + H.  相似文献   

20.
《Applied Thermal Engineering》2007,27(5-6):869-876
In this work, the effects of pore sizes of silica gel on desorption activation energy and adsorption kinetics of water vapour on the silica gels were studied. The isotherms and adsorption kinetic curves of water vapour on three kinds of silica gels with average pore diameters of 2.0 nm, 5.28 nm and 10.65 nm, respectively, were measured by the method of static adsorption, the desorption activation energies of water vapour on silica gels were estimated by using the TPD technique, and the effects of pore sizes of silica gel on adsorption kinetics and desorption activation energy were discussed. Results showed that the isotherm of water vapour on the A-type silica gel with the average pore diameter of 2 nm was of type I, which can be well described by the Langmuir model; the isotherms of water vapour on the B-type and the C-type mesoporous silica gels with the average pore diameters of 5.28 nm and 10.65 nm, respectively, were of type V; and at lower RH, the smaller the average pore size of the silica gel was, the smaller the adsorption rate constant was due to the diffusion resistance in the pores of the silica gels, while at a higher RH, the smaller the average pore size of the silica gel was, the larger the adsorption rate constant was. The desorption activation energies of water on the A-type, the B-type and the C-type silica gels were respectively 35.54 kJ/mol, 31.41 kJ/mol and 26.16 kJ/mol, which suggested that the desorption activation energy of water on the silica gels increased as the pore sizes of the silica gels decreased.  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