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 607 毫秒
1.
Using a previously developed experimental technique, the behavior of small methane and propane hydrate samples formed from water droplets between 0.25 and 2.5 mm in size has been studied in the pressure–temperature area between the ice–hydrate–gas equilibrium line and the supercooled water–hydrate–gas metastable equilibrium line, where ice is a stable phase. The unusual persistence of the hydrates within the area bounded by these lines and the isotherms at T=253 K for methane hydrate or at T=263 K for propane hydrates was observed. This behavior has not previously been reported. For example, in the experiment carried out at 1.9 MPa and 268 K, the methane hydrates existed in a metastable state (the equilibrium pressure at 268 K is 2.17 MPa) for 2 weeks, then immediately dissociated into liquid supercooled water and gas after the pressure was isothermally decreased slightly below the supercooled water–hydrate–gas metastable equilibrium pressure. It was found that dissociation of metastable hydrate into supercooled water and gas was reversible. The lateral hydrate film growth rates of metastable methane and propane hydrates on the surface of supercooled water at a pressure below the ice–hydrate–gas equilibrium pressure were measured. The temperature range within which supercooled water formed during hydrate dissociation can exist and a role of supercooled water in hydrate self-preservation is discussed.  相似文献   

2.
An in-house custom made high pressure adsorption/desorption unit has been designed and fabricated to study reversible hydrogen (H2) intake capacity, hysteresis, kinetics, plateau pressure of various nanomaterials, zeolites and metallic compounds, in the pressure range of 1  P  150 atm. The unit has been used to estimate H2 intake capacity of carbon nanofibers prepared by flame synthesis in the absence of catalyst. H2 adsorption studies have been carried out in the pressure range of 25–100 atm at 297 K. The maximum H2 intake capacity has been observed to be 3.7 wt% at 100 atm.  相似文献   

3.
The effect of a range of operation variables such as pressure, low temperature and H2/CO molar feed ration the catalytic performance of 80%Co/20%Ni/30 wt% La2O3/1 wt% Cs catalyst was investigated. It was found that the optimum operating conditions is a H2/CO = 2/1 molar feed ratio at 260 °C temperature and 2 bar pressure. Reaction rate equations were derived on the basis of the Langmuir–Hinshelwood–Hougen–Watson (LHHW) type models for the Fischer–Tropsch reactions. The activation energy obtained was 59.69 kJ/mol for optimal kinetic model.  相似文献   

4.
The catalytic growth of structured carbon from a C2H4 and C2HCl3 feed promoted by Ni/SiO2 in the presence of H2 over the temperature range 673 K  T  1023 K has been examined. The supported Ni phase exhibited an exclusive cubic symmetry (XRD analysis) with a range of Ni particle sizes (TEM analysis) and a net shift in the distribution to larger particles with increasing reduction temperature (from 20 to 36 nm), accompanied by a decrease in H2 chemisorption. Conversion of C2H4 generated hydrogenation (C2H6), hydrogenolysis (CH4) and decomposition (C + H2) products. Ethane formation was favoured at lower temperatures with C formation increasingly preferred at higher temperatures so that C2H4 decomposition was the predominant process at T > 723 K; significant CH4 production was only observed at T > 900 K. Carbon yield from C2H4 passed through a maximum at 773 K and took the form of high aspect ratio graphitic nanofibres with a central hollow core and diameters in the range 5–180 nm. The carbonaceous product has been characterized by a combination of TEM-EDX, SEM, XRD, BET area and temperature programmed oxidation (TPO). Carbon formation from C2HCl3 exceeded (by a factor of up to an order of magnitude) that generated via the decomposition of C2H4 at the same inlet C:Ni ratio to deliver essentially a carbon yield invariance (9.1 ± 0.3 gC gNi?1) where 898 K  T  1023 K, which represents a carbon efficiency (fraction of carbon in the inlet feed that is converted to a solid carbon product) in excess of 96%. Ni/SiO2 promoted a composite dehydrochlorination/decomposition of C2HCl3 to HCl + C. The nature of the carbon product generated from C2HCl3 is strongly temperature dependent with a shift from a pseudo-fibrous product at 773 K to a predominant nanosphere formation at 923 K. These nanospheres exhibit a wide diameter range (40–700 nm), a significant Cl content (1.1–2.6%, w/w) and a conglomeration or clustering to give a less ordered carbonaceous product than that generated at the lower temperature (773 K). A tentative carbon growth rationale is presented to account for the observed dependence of carbon structure on carbon-containing precursor and reaction temperature.  相似文献   

5.
The reactivity of methane with lattice oxygen of cerium niobate, CeNbO4+δ, was studied by temperature-programmed reduction (TPR) in dry CH4 flow at 523–1073 K. Phase transformations and reduction of cerium niobate at 900–1023 K lead to a massive release of hyperstoichiometric oxygen, in amounts determined by the intermediate-temperature phase composition dependent on thermal history. In this temperature range, CH4–TPR shows prevailing formation of carbon monoxide and steam, suggesting that the synthesis gas generation occurs in parallel with extensive oxidation of H2 on the cerium niobate surface. At 1073 K when δ  0, the reaction of methane with CeNbO4+δ selectively yields synthesis gas with H2/CO ratio close to two.  相似文献   

6.
The hydrogenation of ethyl butyrate, n-butyric acid, and n-butyraldehyde to their corresponding alcohol(s) has been studied over a γ-Al2O3-supported cobalt catalyst using a high-pressure fixed-bed reactor in the temperature range of 473–493 K. H2–D2–H2 switching experiments show that ethyl butyrate and n-butyric acid follow an inverse kinetic isotope effect (KIE) (i.e. rH/rD = 0.50–0.54), whereas n-butyraldehyde did not display any KIE (i.e. rH/rD = 0.98). DRIFTS experiments were performed over the support and catalyst to monitor the surface species formed during the adsorption of ethyl butyrate and n-butyric acid at atmospheric pressure and the desired temperature. Butanoate and butanoyl species are the stable surface intermediates formed during hydrogenation of ethyl butyrate. Hydrogenation of butanoate to a partially hydrogenated intermediate is likely involved in the rate-determining step of ethyl butyrate and butyric acid hydrogenation.  相似文献   

7.
《Ceramics International》2015,41(8):9686-9691
A novel solid state reaction was adopted to prepare Sm0.2Ce0.8O1.9 (SDC) powder. A mixed oxalate Sm0.2Ce0.8(C2O4)1.5·2H2O was synthesized by milling a mixture of cerium acetate hydrate, samarium acetate hydrate, and oxalic acid for 5 h at room temperature. An ultra-fine SDC powder with the primary particle size of 5.5 nm was obtained at 300 °C. The ultra-low temperature for the formation of SDC phase was due to the atomic level mixture of the Sm3+ and Ce4+ ions. The crystal sizes of SDC powders at 300 °C, 550 °C, 800 °C, and 1050 °C were 5.5 nm, 11.4 nm, 24.1 nm and 37.5 nm, respectively. The sintering curves showed that the powder calcined at lower temperature was easier to be sintered owning to its smaller particle size. A solid oxide electrolytic cell (SOEC), comprising porous La0.8Sr0.2Cu0.1Fe0.9O3−δ (LSCF) for substrate, LSCF–SDC for active electrode, SDC for electrolyte, and LSCF–SDC for symmetric electrode, was fabricated by dip-coating and co-sintering techniques. An extremely dense SDC film with the thickness of 20 μm was obtained at only 1200 °C, which was about 100–300 °C lower than the literatures׳ reports. The designed SOEC was proved to work effectively for decomposing NO (3500 ppm, balanced in N2), 80% NO can be decomposed at 600 °C.  相似文献   

8.
The influence of sintering temperature, holding time and pressure condition on densification and mechanical properties of bulk titanium carbide (TiC) fabricated by SPS sintering has been systematically investigated. Experimental data demonstrated that relative density and Vickers hardness (HV) increase with sintering temperature and holding time, but fracture toughness (KIC) was not significantly influenced by sintering parameters. The HV and relative density of samples consolidated by SPS technique at 1600 °C for 5 min under 50 MPa pressure (applied entire sintering cycle) reached 30.31 ± 2.23 GPa and 99.90%, respectively. HV values of ~24–30 GPa and KIC of ~3.7–5 MPa m1/2 were obtained in all bulk samples with relative densities of 95.61–99.90% when fabricated under various conditions presented above, without abnormal grain growth. More pronounced effects of pressure condition on grain growth (promoted by grain-boundary diffusion) than on densification were observed. The relationship of fracture toughness and fracture mode is also discussed.  相似文献   

9.
Twelve strains of Clostridium botulinum group I spores, suspended in phosphate buffer (0.1 M) at approximately 107 CFU/ml concentration, were subjected to high pressure treatments (800 and 900 MPa; 0.5–15 min) at elevated temperatures (90 and 100 °C). The treatments were chosen to have a range of pressure/temperature severity to be able to discriminate the spore strains for their pressure resistance. An insulated test chamber was used to achieve temperature stability during treatment. Preliminary tests showed the need for an 8 day anaerobic incubation for enumeration. Results showed that strains PA9508B, HO9504A and CK2-A had higher pressure resistance than others among the 12 strains studied. Strain 62A was least resistant and completely inactivated by the treatment. Estimated D values of the more resistant strains were in the 0.66–1.8 min range at 900 MPa at 100 °C treatment. The temperature sensitivity parameter (ZP value) in the 800–900 MPa pressure range varied between 10 and 16 °C, and pressure sensitivity parameter (ZT value) in the 90–100 °C temperature range varied between 340 and 760 MPa. The most pressure resistant strain was PA9508B with an estimated ZP value of 16.0 °C and ZT value of 470 MPa. Since the pathogenic strains of C. botulinum have different pressure resistance, the most resistance strain should be selected as the basis for process establishment.  相似文献   

10.
The high-pressure vapour–liquid phase equilibria (PTxy) of the binary mixture propylene glycol/CO2 have been experimentally investigated at temperatures of (398.2, 423.2 and 453.2) K over the pressure range from (2.5 to 55.0) MPa using a static-analytic method. Furthermore, the high-pressure vapour–liquid phase equilibria (PTxy) of the ternary mixture propylene glycol/CO2/ethanol at constant temperatures of (398.2, 423.2 and 453.2) K and at constant pressure of 15.0 MPa have been determined using a static-analytic method. Initial concentrations of components in propylene glycol (PG)/ethanol (EtOH) mixture vary from 10 up to 90 wt.%. In general, for binary system it was observed that the solubility of CO2 in the heavy propylene glycol reach phase increases with increasing pressure at constant temperature. On the contrary, the composition of gaseous phase is not influenced by the pressure or the temperature. On average the solubility of PG in light phase of CO2 amounts to 30 wt.%. The system behaviour at temperature of 398.2 K was investigated up to 70.0 MPa and a single-phase region was not observed. Above the pressure 60.0 MPa a single-phase region of the system was observed for the temperature of 423.2 K. For the temperature of 453.2 K the single-phase was observed above the pressure of 48.0 MPa. For ternary system it was observed that the composition of heavy phase is slightly influenced by the temperature when the mass fraction of EtOH in initial mixture is higher than 50 wt.%. If the mass fraction of PG in initial mixture is higher than 50 wt.%, the composition of heavy phase is not influenced by the temperature anymore. The composition of the PG, EtOH and CO2 in light phase remains more or less unchanged and it is not influenced by the conditions.  相似文献   

11.
Broadband dielectric spectroscopy results of various ordered and disordered (1 ? x)Pb(Mg1/3Nb2/3)O3–(x)Pb(Sc1/2Nb1/2)O3 (PMN–PSN) ceramics are investigated in the temperature range from 80 K to 300 K and frequency range from 20 Hz to 2 THz. Dielectric dispersion is very broad and in the ferroelectrics case (x = 1, 0.95) consists of two parts: low-frequency part caused by ferroelectric domains and higher frequency part caused by soft mode. The relaxational soft mode exhibits pronounced softening close to phase transition temperature, as it is typical for order–disorder phase transitions. By substituting Sc3+ by Mg2+ in PMN–PSN ceramics relaxation slows down, and for relaxors (x = 0.2) the most probable relaxation frequency decreases on cooling according to Vogel–Fulcher law.  相似文献   

12.
The phase behavior of hexamethyldisiloxane (HMDS)–carbon dioxide (CO2) binary mixture was investigated using a constant volume view cell. The accuracy of the measurement technique was inspected against the bubble point pressure data in the literature for ethanol (C2H5OH)–carbon dioxide (CO2) binary mixture. The bubble point pressures for C2H5OH–CO2 agreed well with the literature values. The bubble point pressures of HMDS–CO2 binary mixture were determined at five different temperatures (T = 298.2 K, 308.2 K, 313.2 K, 323.2 K, 333.2 K) and at various compositions. The bubble point pressures increased with increasing temperature and CO2 mole fraction in the binary mixture. The phase behavior of the binary mixture was modeled using the Peng–Robinson Stryjek–Vera equation of state (PRSVEoS). The binary interaction parameters were regressed from experimental bubble point pressures at each temperature and were found to exhibit a linear dependency on temperature. The HMDS–CO2 binary mixture was also found to exhibit Type II phase behavior. Additionally, PTρ measurements for the same binary system were conducted and excess molar volumes were calculated.  相似文献   

13.
The gasification reactivity of char from dried sewage sludge (DSS) applicable to fluidized bed gasification (FBG) was determined. The char was generated by devolatilizing the DSS with nitrogen at the selected bed temperature and was subsequently gasified by switching the fluidization agent to mixtures of CO2 and N2 (CO2 reactivity tests) and steam and N2 (H2O reactivity tests).. The tests were conducted in the temperature range of 800–900 °C at atmospheric pressure, using partial pressure of the main reactant in the mixture (CO2 or H2O) in the range of 0.10–0.30 bar. Expressions for the intrinsic reactivity (free of diffusion effects) as a function of temperature, partial pressure of gas reactant (CO2 or H2O) and degree of conversion were obtained for each reaction. For the whole range of conversion it was found that the char reactivity in an H2O–N2 mixture was roughly three times higher than that in a mixture with the corresponding partial pressure of CO2. The reactivity was only influenced by particle size greater than 1.2 mm in the tests with steam at 900 °C. It was demonstrated that the method of char preparation greatly influences the reactivity, highlighting the importance of generating the char in conditions similar to that in FBG.  相似文献   

14.
Six-carbon containing model compounds of 1-hexanol, 1,6-hexanediol and sorbitol for the reforming were employed to study the effect of hydroxyl functional group on the thermodynamic equilibrium product distribution in a wide range of conditions of temperature (300–1300 K), pressure (1–150 atm) and feed composition (H2O/carbon = 0.5–30). The increase of hydroxyl group in reactants gave rise to the increase of carbon dioxide with loss of methane in the product in the following order: sorbitol > 1,6-hexanediol > 1-hexanol, while the formation of hydrogen and carbon monoxide was mostly governed by the feed composition (H2O/carbon) and pressure rather than the number of hydroxyl group in reactants.  相似文献   

15.
Recently, trivalent rare earth doped materials have received significant attention due to the strong temperature dependence of the fluorescence emission of these materials, which can be useful in temperature sensing. Here, we investigated Y2O3 ceramic powders doped with Yb3+ and co-doped with either Tm3+ or Ho3+. The powders were obtained via spray pyrolysis at 900 °C and additionally thermally treated at 1100 °C for 24 h. Structural characterization using X-ray powder diffraction confirmed the cubic bixbyte structure. Scanning electron microscopy (SEM) revealed that the particles exhibit a uniform spherical morphology. The up-conversion emissions were measured using laser excitation at 978 nm, resulting in the following transitions: blue emission in the range of 450–500 nm, weak red emission in the range of 650–680 nm and near infrared emission in the range of 765–840 nm for Tm3+, as well as green emission centered at 550 nm and weak near infrared emission at 755 nm for the Ho3+ ions. In addition, the temperature dependence of the fluorescence intensity ratios of different Stark components was analyzed in the range of 10–300 K. Significant temperature sensitivity was detected for several components, with the largest value of 0.097 K?1 related to the intensity ratio of I536 and I772 emissions observed for the Y2O3:Yb,Ho powder.  相似文献   

16.
New high temperature negative temperature coefficient (NTC) thermistor ceramics based on a xMgAl2O4–(1  x)YCr0.5Mn0.5O3 (x = 0.1, 0.4, 0.6) composite system have been successfully fabricated through spark plasma sintering (SPS) with a low sintering temperature and a short sintering period. The X-ray diffraction analysis indicates that the SPS-sintered composite ceramics consist of a cubic spinel MgAl2O4 phase and an orthorhombic perovskite YCr0.5Mn0.5O3 phase isomorphic to YCrO3. The SPS-sintered composite ceramics have high relative density ranging from 94.1 to 97.4% of the theoretical density. X-ray photoelectron spectroscopy analysis corroborates the presence of Cr3+, Cr4+, Mn3+, and Mn4+ ions on lattice sites, which may result in the hopping conduction. The obtained ρ25, B25–150, and B700–1000 of the SPS-sintered composite NTC thermistors are in the range of 1.53 × 106–9.92 × 109 Ω cm, 3380–5172 K, and 7239–9543 K, respectively. These values can be tuned by adjusting the MgAl2O4 concentration.  相似文献   

17.
Mn2+-doped Sn1−xMnxP2O7 (x = 0–0.2) are synthesized by a new co-precipitation method using tin(II)oxalate as tin(IV) precursor, which gives pure tin pyrophosphate at 300 °C, as all the reaction by-products are vaporizable at <150 °C. The dopant Mn2+ acts as a sintering aid and leads to dense Sn1−xMnxP2O7 samples on sintering at 1100 °C. Though conductivity of Sn1−xMnxP2O7 samples in the ambient atmosphere is 10−9–10−6 S cm−1 in 300–550 °C range, it increases significantly in humidified (water vapor pressure, pH2O = 0.12 atm) atmosphere and reaches >10−3 S cm−1 in 100–200 °C range. The maximum conductivity is shown by Sn0.88Mn0.12P2O7 with 9.79 × 10−6 S cm−1 at 550 °C in ambient air and 2.29 × 10−3 S cm−1 at 190 °C in humidified air. It is observed that the humidification of Sn1−xMnxP2O7 samples is a slow process and its rate increases at higher temperature. The stability of Sn1−xMnxP2O7 samples is analyzed.  相似文献   

18.
Microwave dielectric properties of CaWO4 ceramics were investigated as a function of H3BO3 and/or Bi2O3 content and sintering temperature. For a single addition of H3BO3 (1  x (wt.%)  5), the density of specimen increased up to 3 wt.% H3BO3, and then decreased. The dielectric constant (K) and the quality factor (Q × f) of the specimens sintered at 850 °C showed lower value than those of specimens sintered above 900 °C due to the poor sinterability. With the increase of H3BO3 content of 0.5 wt.% Bi2O3yH3BO3 (5  y (wt.%)  20), the sintering temperature of CaWO4 ceramics could be effectively reduced from 1100 to 850 °C without degradation of dielectric properties. For the specimens sintered at 850 °C for 30 min, K was not changed remarkably with Bi2O3–H3BO3 content; however, Q × f value increased up to 9 wt.% H3BO3 of 0.5 wt.% Bi2O3yH3BO3, and then decreased. The temperature coefficient of resonant frequency (TCF) shifted to the positive value with increasing Bi2O3–H3BO3 content. Typically, K of 8.7, Q × f of 70,220 GHz and TCF of −15 ppm/°C were obtained for the specimens with 0.5 wt.% Bi2O3–9 wt.% H3BO3 sintered at 850 °C for 30 min.  相似文献   

19.
Perovskites La1−xCaxAlyFe1−yO3−δ (x, y = 0 to 1) were prepared by high-temperature solid-state synthesis based on mixtures of oxides produced by colloidal milling. The XRD analysis showed that perovskites La0.5Ca0.5AlyFe1−yO3−δ with a high Fe content (1  y = 0.8–1.0) were of orthorhombic structure, perovskites with a medium Fe content (1  y = 0.8–0.5) were of rhombohedral structure, and perovskite with the lowest Fe content (1  y = 0.2) were of cubic structure. Thermally programmed desorption (TPD) of oxygen revealed that chemical desorption of oxygen in the temperature range from 200 to 1000 °C had proceeded in the two desorption peaks. The low-temperature α-peak (in the 200–550 °C temperature range) was brought about by oxygen liberated from oxygen vacancies; the high-temperature β-peak (in the 550–1000 °C temperature range) corresponded to the reduction of Fe4+ to Fe3+. The chemidesorption oxygen capacity increased with increasing Ca content and decreased with increasing Al content in the perovskites. The Al3+ ions restricted, probably for kinetic reasons, the reduction of Fe4+ and the high-temperature oxygen desorption associated with it.  相似文献   

20.
A set of porous carbons has been prepared by chemical activation of various fungi-based chars with KOH. The resulting carbon materials have high surface areas (1600–2500 m2/g) and pore volumes (0.80–1.56 cm3/g), regardless of the char precursors. The porosities mainly derived from micropores in activated carbons strongly depend on the activation parameters (temperature and KOH amount). All activated carbons have uniform micropores with pore size of 0.8–0.9 nm, but some have a second set of micropores (1.3–1.4 nm pore size), further broadened to 1.9–2.1 nm as a result of increasing either the activation temperature to 750 °C or KOH/char mass ratio to 5/1. These fungi-based porous carbons achieve an excellent H2 uptake of up to 2.4 wt% at 1 bar and −196 °C, being in agreement with results from other porous carbonaceous adsorbents reported in the literature. At high pressure (ca. 35 bar), the saturated H2 uptake reaches 4.2–4.7 wt% at −196 °C for these fungi-based porous carbons. The results imply a great potential of these fungi-based porous carbons as H2 on-board storage media.  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