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 375 毫秒
1.
Abstract

Hydroxyapatite (HAp) with the general formula Ca10 z(HPO4)z (PO4) 6-z (OH) a-z · nH2O has become of particular interest as a bio-restorative material having good compatibility with living hard tissues. In this work, hydration and hardening properties of water-setting apatitic cements were presented. Ca3(PO4)2 with the formd α (α-TCP) itself was an apatitic cement which could hydrate and harden alone, and formed HAp or Ca8H2 (PO4) 6 · 5H2O (OCP). Other apatitic cements were prepared by mixing at least two compounds selected from Ca HPO4 · 2H2O (DCPD), α-TCP, Ca4(PO4)aO (TeCP) and CaCO3. The hydraulic reaction of α-TCP was accelerated by using water-soluble additives such as NaCl, NH4Cl, NH4 -citrate, etc. Another way for the acceleration was the addition of DCPD. The combination α-TCP/DCPD was developed with the expectation of the reaction α-TCP + DCPD + H2O → 1/2 OCP. Hydraulic reactions of combinations TeCP/DCPD (Brown and Chow, 1983) and DCPD/CaCO3 were also investigated. Various setting times above 4 min were obtained at 37–40°C by using different additives and combinations. The setting and hardening was considered to be due to the entanglement of HAp or OCP microcrystals formed on original powder particles. The resulting porously hardened bodies had porosities of 46 –80%, compressive strenghts of 2–30MPa, diametral tensile strengths of 0.1–3.5MPa. These apatitic cement pastes showed different DSC characteristics;d-TCP: exothexmic peak at 70–105°C, α-TCP/DCPD:two exo. and two endothermic peaks in the range of 40 to 120°C, TeCP/DCPD:two exothermic peaks in the range of 30 to 80°C, and DCPD/CaCO3:endothermic peak at 70–100 °C.  相似文献   

2.
The living cationic polymerization of 6‐tert‐butoxy‐2‐vinylnaphthalene (tBOVN), a vinylnaphthalene derivative with an electron‐donating group, was achieved with a TiCl4/SnCl4 combined initiating system in the presence of ethyl acetate as an added base at –30 °C. The absence of side reactions at low temperature was confirmed by 1H NMR analysis of the resulting polymer. In contrast to this controlled reaction at –30 °C, reactions performed at higher temperature, such as 0 °C, frequently involved unwanted intramolecular or intermolecular Friedel–Crafts reactions of naphthalene rings due to the high electron density of these rings. The cationic polymerization of 6‐acetoxy‐2‐vinylnaphthalene, a derivative with an acetoxy group, was also controlled under similar conditions, but chain transfer reactions were not completely suppressed during the polymerization of 2‐vinylnaphthalene. The glass transition temperature (Tg) of the obtained poly(tBOVN) was 157 °C, a value higher by 94 °C than that of the corresponding styrene derivative. © 2013 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2013 , 51, 4828–4834  相似文献   

3.
Reactions of ozone with simple olefins have been studied between 6 and 800 mtorr total pressure in a 220-m3 reactor. Rate constants for the removal of ozone by an excess of olefin in the presence of 150 mtorr oxygen were determined over the temperature range 280 to 360° K by continuous optical absorption measurements at 2537 Å. The technique was tested by measuring the rate constants k1 and k2 of the reactions (1) NO + O3 → NO2 + O2 and (2) NO2 + O3 rarr; NO3 + O2 which are known from the literature. The results for NO, NO2, C2H4, C3H6, 2-butene (mixture of the isomers), 1,3→butadiene, isobutene, and 1,1 -difluoro-ethylene are 1.7 × 10?1 4 (290°K), 3.24 × 10?17 (289°K), 1.2 × 10?1 4 exp (–4.95 ± 0.20/RT), 1.1 × 10?1 4 exp (–3.91 ± 0.20/RT), 0.94 × 10?1 4 exp ( –2.28 ± 0.15/RT), 5.45 ± 10?1 4 exp ( –5.33 ± 0.20/RT), 1.8 ×10?17 (283°K), and 8 × 10?20 cm3/molecule ·s(290°K). Productformation from the ozone–propylene reaction was studied by a mass spectrometric technique. The stoichiometry of the reaction is near unity in the presence of molecular oxygen.  相似文献   

4.
Low-temperature steam conversion (LTSC) of a methane-butane mixture (95% methane and 5% butane) into a methane-rich gas over an industrial Ni-based catalyst has been studied with the following reaction conditions: temperature 200–320°C, pressure 1 bar, gas hour space velocity 1200–3600 h–1, and steam to carbon ratio 0.64. A three-step macrokinetic model has been suggested based on the kinetic parameters found. The model includes the following reactions: (1) irreversible steam reforming; (2) CO2 methanation, which occurs in a quasi-equilibrium mode at temperatures above 260°C; (3) hydrogenolysis of propane and butane, which is essential at temperatures below 260°C. Steam reforming was shown to limit the overall reaction rate, whereas hydrogenolysis and CO2 methanation determined the product distribution in low- and high-temperature regions, respectively. Temperature dependencies of the product distribution for the LTSC of a model ternary methane-propane-butane mixture (85% methane, 10% propane, and 5% butane) have been successfully simulated using the three-step model suggested.  相似文献   

5.
Triclinic single crystals of Cu4(H3N–(CH2)9–NH3)(OH)2[C6H2(COO)4]2 · 5H2O were prepared in aqueous solution at 80 °C in the presence of 1,9‐diaminononane. Space group P$\bar{1}$ (no. 2) with a = 1057.5(2), b = 1166.0(2), c = 1576.7(2) pm, α = 106.080(10)°, β = 90.73(2)° and γ = 94.050(10)°. The four crystallographic independent Cu2+ ions are surrounded by five oxygen atoms each with Cu–O distances between 191.4(3) and 231.7(4) pm. The connection between the Cu2+ coordination polyhedra and the [C6H2(COO)4]4– anions yields three‐dimensional framework with negative excess charge and wide centrosymmetric channel‐like voids. These voids extend parallel to [001] with the diagonal of the nearly rectangular cross‐section of approximately 900 pm. The channels of the framework accommodate [H3N–(CH2)9–NH3]2+ cations and water molecules, which are not connected to Cu2+. The nonane‐1,9‐diammonium cations adopt a partial gauche conformation. Thermoanalytical measurements in air show a loss of water of crystallization starting at 90 °C and finishing at approx. 170 °C. The dehydrated compound is stable up to 260 °C followed by an exothermic decomposition yielding copper oxide.  相似文献   

6.
The title compounds were studied by TG, DTA, DSC, IR and absorption spectroscopy. The complexes go through dehydration (70–200°C), an irreversible exothermic process (in air or N2 atmosphere, 250–300°C) and decomposition to a mixture of oxides and carbonates (385–700°C). The exothermic process occurs without weight loss and corresponding heats of reaction fall in the range 0–26 kJ/mol. The absorption spectrum of the Nd complex in the range 5000–6000 Å was employed to monitor perturbations in the coordination sphere of Nd3+ arising from the exothermic process. Involvement of the Nd3+ cation is implied and the heats of reaction show a close relationship to the radii of Ln3+. The interpretation of these data was made with the aid of valuable structural information obtained previously.  相似文献   

7.
Density, Conductivity, and Electrolysis of Liquid Phases in Nonaqueous Systems of the Type MCl/AlCl3/SO2 (M = Li, Na) The temperature dependence of the density and specific conductivity was determined at liquid phases of the composition MAlCl4 · nSO2 + mAlCl3 (M = Li, n = 3–5.5, m = 0.266; M = Na, n = 1.36–4.56, m = 0.01–0.1). The investigated range was between ?30°C and +45°C. For different compositions it was limited by the liquidus point and by the point, where the SO2-equilibrium pressure surpassed 1 bar. The densities are between 1.63 and 1.76 g/cm3, the specific conductivities between 0.03 and 0.07 Ω?1 · cm?1. In diluted solutions below ?10°C NaAlCl4 behaves as a strong electrolyte in which dissociation in Na+ and AlCl4?is prevailing. By electrolysis of the liquid phases at room temperature reversible galvanic cells of the type M/MAlCl4 · nSO2/Cl2, C(M = Li, Na) are generated, which have an open circuit potential between 4.12 and 4.18 volts. The alkali metal deposits are stable in contact with the electrolyte up to 60°C in the case of lithium and 35°C with sodium.  相似文献   

8.
Thermogravimetric-mass spectrometric (TG/MS) and differential scanning calorimetric (DSC) techniques were used in the characterization of oxidative and nonoxidative degradation reactions of a highly crosslinked divinylbenzene/styrene copolymer. When the copolymer was subjected to a temperature-programmed air environment, four exothermic reactions were detected. The initial small exothermic reaction, starting at ca. 125°C and reaching its maximum at ca. 180°C, was presumed to result from the decomposition of peroxides. The second exothermic reaction, which overlapped with the initial one and peaked at ca. 270°C, was attributed to oxidation with a significant amount of oxygen uptake and liberation of some gaseous products such as CO2, styrene, benzaldehyde, ethylstyrene, and ethylbenzaldehyde. The strongest exothermic reaction took place at ca. 290–380°C and had its peak at ca. 360°C. Associated with this reaction was the generation of many gaseous pyrolysates, as given above. The exothermic reaction continued at a relatively constant rate from ca. 380°C to the maximum temperature of the experiment (500°C) with the release of only one gaseous product (CO2). The initial exothermic reaction can be eliminated by controlled thermal decomposition of peroxides; therefore, a more thermally stable polymer can be obtained. Exothermic reactions, starting at ca. 170°C, were observed. Pyrolytic reactions in an inert gas were also studied.  相似文献   

9.
High pressure DSC and simultaneous TG/DSC were used to study the different kinds of boron that was used in the fuel-rich propellant and the amorphous boron in different gases and different pressure. Also, some of the samples before the experiment and after the experiment were analyzed by the SEM. The results show that: (1) there is a big exothermic peak between 550 °C and 850 °C for all the samples because the combustion heat of boron is very high, and the exothermic peak appears in advance when the pressure or the oxygen concentration increases. (2) Although the reaction process of all the samples with air or oxygen could be divided into five stages, the reaction characteristics are different from each other. Especially, amorphous boron is much more active than the boron used in the fuel-rich propellant. (3) The exothermic peak at about 700 °C appears in advance, and the percentage conversion of boron decreases when the content of magnesium increases and boron–magnesium compound is used as the raw materials. (4) Some samples start to lose their mass for the sake of the evaporation of the (BO)n.  相似文献   

10.
One of the rare alternative reagents for the reduction of carbon–carbon double bonds is diimide (HN?NH), which can be generated in situ from hydrazine hydrate (N2H4 ? H2O) and O2. Although this selective method is extremely clean and powerful, it is rarely used, as the rate‐determining oxidation of hydrazine in the absence of a catalyst is relatively slow using conventional batch protocols. A continuous high‐temperature/high‐pressure methodology dramatically enhances the initial oxidation step, at the same time allowing for a safe and scalable processing of the hazardous reaction mixture. Simple alkenes can be selectively reduced within 10–20 min at 100–120 °C and 20 bar O2 pressure. The development of a multi‐injection reactor platform for the periodic addition of N2H4 ? H2O enables the reduction of less reactive olefins even at lower reaction temperatures. This concept was utilized for the highly selective reduction of artemisinic acid to dihydroartemisinic acid, the precursor molecule for the semisynthesis of the antimalarial drug artemisinin. The industrially relevant reduction was achieved by using four consecutive liquid feeds (of N2H4 ? H2O) and residence time units resulting in a highly selective reduction within approximately 40 min at 60 °C and 20 bar O2 pressure, providing dihydroartemisinic acid in ≥93 % yield and ≥95 % selectivity.  相似文献   

11.
《Mendeleev Communications》2023,33(4):574-576
The ignition temperatures and effective activation energies of the ignition limits of mixtures (40–70% H2 + 60–30% CH4)stoich + air over Rh were experimentally determined at a pressure of 1 atm in the temperature range 20–300 °C. Over an ignition-treated Rh surface, the ignition temperature of a mixture of 70% H2 + 30% methane + air is 62 °C. This indicates the potential of using Rh to markedly lower the ignition temperature of fuels based on hydrogen–methane mixtures.  相似文献   

12.
The aim of current work is to study the interaction of process parameters including, temperature, CO2 feed composition and feed pressure were towards CO2 separation from CO2/CH4 binary gas mixture over hollow fiber mixed matrix membrane using design of experiment (DoE) approach. The hollow fiber mixed matrix membrane (HFMMM) containing NH2-MIL-53(Al) filler and cellulose acetate polymer was successfully spun and fibers with outer diameter of approximately 250–290 nm were obtained. The separation results revealed that the increment of temperature from 30 °C to 50 °C reduced the CO2/CH4 separation factor while, increasing feed pressure from 3 bar to 15 and increment of CO2 feed composition from 15 to 42.5 vol% increased the separation factor of HFMMM. The DoE results showed that the feed pressure was the most significant process parameter that intensely affected the CH4 permeance, CO2 permeance and CO2/CH4 separation factor. Based on the experimental results obtained, maximum CO2 permeance of 3.82 GPU was achieved at feed pressure of 3 bar, temperature of 50 °C and CO2 feed composition of 70 vol%. Meanwhile, minimum CH4 permeance of 0.01 GPU was obtained at feed pressure of 15 bar and temperature of 30 °C and CO2 feed composition of 70 vol%. Besides, maximum CO2/CH4 separation factor of 14.4 was achieved at feed pressure of 15 bar and temperature of 30 °C and CO2 feed composition of 15 vol%. Overall, the study on the interaction between separation processes parameters using central composite design (CCD) coupled with response surface methodology (RSM) possesses significant importance prior to the application of NH2-MIL-53(Al)/Cellulose Acetate HFMMM at industrial scale of natural gas purification.  相似文献   

13.
《Fluid Phase Equilibria》1999,157(2):285-297
Cloud-point data for the system poly(methyl methacrylate) (PMMA)–CO2–methyl methacrylate (MMA) are measured in the temperature range of 26 to 170°C, to pressures as high as 2500 bar, and with cosolvent concentrations of 10.4, 28.9, and 48.4 wt.%. PMMA does not dissolve in pure CO2 to 255°C and 2550 bar. The cloud-point curve for the PMMA–CO2–10.4 wt.% MMA system exhibits a negative slope that reaches 2500 bar at 105°C. With 28.9 wt.% MMA the cloud-point curve remains relatively flat at ∼900 bar for temperatures between 25 and 170°C. With 48.4 wt.% MMA the cloud-point curve exhibits a positive slope that extends to 20°C and ∼100 bar. Pressure-composition isotherms are also reported for the CO2–MMA system at 40.0, 80.0, 105.5°C. This system exhibits type-I phase behavior with a continuous mixture–critical curve. The Peng–Robinson (PR) and SAFT equations of state model the CO2–MMA data reasonably well without any binary interaction parameters, although the PR equation provides a better representation of the mixture-critical region. It is not possible to obtain even a qualitative fit of the PMMA–MMA–CO2 data with the SAFT equation of state. The SAFT model qualitatively shows that the cloud-point pressure decreases with increasing MMA concentration and that the cloud-point curve exhibits a positive slope for very high concentrations of MMA in solution.  相似文献   

14.
Diluted (1.0–1.5 wt%) aqueous gelatin solutions have been cooled to –10 °C at a cooling rate 20 °C min−1 without freezing and detectable gelation. When heated at a constant heating rate (0.5 –2 °C min−1), the obtained supercooled solutions demonstrate an atypical process of gelation that has been characterized by regular and stochastically modulated differential scanning calorimetry (DSC) as well as by isoconversional kinetic analysis. The process is detectable as an exothermic peak in the total heat flow of regular DSC and in the nonreversing heat flow of stochastically modulated DSC. Isoconversional kinetic analysis applied to DSC data reveals that the effective activation energy of the process increases from approximately 75 to 200 kJ mol−1 as a supercooled solution transforms to gel on continuous h eating.  相似文献   

15.
TlPd3 was synthesised from the elements in evacuated silica tubes at 600 °C. Alternatively, TlPd3 was yielded by reduction of TlPd3O4 in N2 gas atmosphere. Reduction of the oxide in H2 gas atmosphere resulted in the formation of the new hydride TlPd3H. The structure of tetragonal TlPd3 (ZrAl3 type, space group I4/mmm, a = 410.659(9) pm, c = 1530.28(4) pm) was reinvestigated by X‐ray and also by neutron powder diffraction as well as the structure of its previously unknown hydride TlPd3H (cubic anti‐perovskite type structure, space group Pm\bar{3} m, a = 406.313(1) pm). In situ DSC measurements of TlPd3 in hydrogen gas atmosphere showed a broad exothermic signal over a wide temperature range with two maxima at 280 °C and at 370 °C, which resulted in the product TlPd3H. A dependency of lattice parameters of the intermetallic phase on reaction conditions is observed and discussed. Results of hydrogenation experiments at room temperature with gas pressures up to 280 bar hydrogen and at elevated temperatures with very low hydrogen gas pressures (1–2 bar) as well as results of dehydrogenation of the hydrides under vacuum will be discussed.  相似文献   

16.
《Mendeleev Communications》2022,32(3):405-407
The flammability limits of stoichiometric mixtures (20–80% H2 + 80–20% CH4) + O2 over Rh and Pd were determined in the pressure range 0–200 Torr and the temperature range 200–500 °C. It has been shown that the dark reaction in the mixture (80% H2 + 20% C4H8)stoich + O2 leads to the formation of carbon nanotubes with a mean diameter of 10–100 nm.  相似文献   

17.
The apparent first ionization constant of carbonic acid has been determined by conductivity measurements and found to vary from 4.32×10?7 at 25°C to 1.6×10?8 at 250°C. The pressure effect to 2000 bar has been measured, and the ratio Kp/K1 is 7.3 at 25°C and 19 at 250°C. The standard partial molar volume change for the ionization at 1 bar, \(\Delta \bar V_1^0\) , increases from ?27.6 cm3-mole?1 at 25°C to ?88 cm3-mole?1 at 250°C. The volume changes are smaller at higher pressures. A linear correlation between \(\Delta \bar V_1^0\) and the partial molar compressibility for the ionization reaction has been noted. A similar correlation exists between the partial molar entropy and volume changes for the reaction.  相似文献   

18.
Commercial polystyrene has been chemically modified with 4,4-dinitro valeryl chloride by use of Friedel–Crafts acylation reaction in the presence of anhydrous aluminum chloride in a mixture of 1,2-dichloroethane and nitrobenzene. The modified polystyrene containing –COCH2CH2C(NO2)2CH3 fragments in side phenyl rings, named gem-dinitro valerylated polystyrene (GDN-PS), was characterized by an Ubbelohde’s viscometer, FTIR, and 1H NMR spectroscopy. Simultaneous thermogravimetry–differential thermal analysis and differential scanning calorimetry (DSC) have been used to study thermal behavior of the polymer. The results of TG analysis revealed that the main thermal degradation for the GDN-PS occurs during two temperature ranges of 200–300 and 300–430 °C. The DTA curve of GDN-PS is showing a visible exothermic peak at 253.8 °C corresponding to the decomposition of gem-dinitro valeryl groups. The decomposition kinetic of the gem-dinitro groups for GDN-PS with degree of substitution (DS) 11 % was studied by non-isothermal DSC under various heating rates. Kinetic parameters such as activation energy and frequency factor for thermal decomposition of GDN-PS with DS 11 % were evaluated via the ASTM E698 and two isoconversional methods.  相似文献   

19.
Reaction of diphenyldiacetylene (1,4-diphenylbutadiyne) by annealing under elevated pressure (0.1–500 MPa) was carried out. Diphenyldiacetylene reacted at 210°C with appearance of gas, and this temperature was independent of the pressure. The measurement of high pressure differential thermal analysis (DTA) revealed that the reaction temperature under elevated pressure was below 24–42°C of the exothermic peak temperature. This implied that exothermic reaction occurred under elevated pressure. Elementary analysis, gel permeation chromatography (GPC), FTIR, Raman scattering, and high resolution 13CNMR experiments were performed to characterize the structure of the product. It was indicated that the product was a mixture of derivative of condensed polycyclic aromatic compound with phenyl group and diphenyldiacetylene oligomer. The fraction of the derivatives increased with increasing pressure, and pressure accelerated the dehydrogenation of the derivatives. The number-averaged molecular weight (Mn) of the diphenyldiacetylene oligomer was 470–610 and the weight-averaged molecular weight (Mw) was 1700–2300. It was considered that the oligomer had a polyacene-based structure. © 1994 John Wiley & Sons, Inc.  相似文献   

20.
A high throughput experimentation (HTE) study identified active Ru metathesis catalysts and reaction conditions for the ethenolysis of maleate esters to the respective acrylate esters. Catalysts were tested at various loadings (75–10’000 ppm) and temperatures (30–60 °C) with maleate esters dissolved in toluene (up to ca. 44 wt-%) or neat and at variable partial pressures of ethylene (0.2–10 bar). Ruthenium catalysts containing a PCy3 ligand, such as 1st or 2nd generation Grubbs catalysts, as well as the state-of-the-art catalysts containing cyclic alkyl amino carbene (CAAC) ligands, are generally inferior to Hoveyda–Grubbs 2nd generation catalyst in ethenolysis of maleates. Productive turnover numbers could exceed 1900 if the ethenolysis reaction is performed at low ethylene pressure (0.2–3 bar) and reach 5200 when a polymeric phenol additive was used. Such catalytic performance falls well within the window practiced in industry. Moreover, a crude technoeconomic analysis finds similar production cost for the ethenolysis route and conventional technology, that is, propene oxidation followed by esterification, justifying research to further improve the ethenolysis route.  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