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 chemical structure of the prepared cationic surfactants which formed through condensation reaction between dimethylaminopropylamine (DMAPA) and butyraldehyde then quaternized by three fatty alkyl bromide was confirmed by FTIR, 1HNMR and mass spectroscopy. The chemical structure of prepared compounds has an effect on surface properties. By increasing the hydrophobic chain length, the values of CMC and Гmax decrease while Amin value was increased. The Thermodynamic parameters showed that adsorption and micellization processes are spontaneous. It is clear that the prepared cationic surfactants tend to adsorb at surface, then it aggregate to form micelle. The prepared surfactants showed good biological activity against Gram-positive and negative bacteria and fungi. The prepared cationic surfactant showed aggressive effect on the sulfate reducing bacteria growth.  相似文献   

2.
A series of amphoteric celluloses (QACMCs) were prepared from sodium carboxymethyl cellulose (CMC) and 2,3-epoxypropyltrimethylammonium chloride (EPTMAC) with a cationic degree of substitution (DScat) of 0.24–1.06 and a carboxymethyl degree of substitution (DSani) of 0.60. The structures of the samples were characterized by FTIR and NMR spectroscopy, which revealed that DScat depended on the ratio of EPTMAC to CMC in the reaction mixture and that the substituent distribution of the cationic group at the C2, C3, and C6 positions of the QACMCs ranked as follows: C2 > C6 > C3. The QACMCs dissolved over a wide range of pH levels and exhibited flocculation ability against kaolin suspensions. In particular, the samples with high DScat exhibited excellent flocculation performance. In addition, the flocculation characteristics of the QACMCs depended on the pH of the kaolin suspension.  相似文献   

3.
Poly(N-vinyl-2-pyrrolidone) (PVP) particles were prepared by dispersion polymerization in the presence of 2,2′-azobisisobutyronitrile as the initiator and siloxane-based surfactant in supercritical carbon dioxide (scCO2). The dispersants used in this study were non-ionic, non-reactive and commercially produced siloxane-based surfactants (Monasil PCA and KF-6017). We investigated the effect of kinds and concentrations of the surfactants, in addition to the reaction temperature and the concentration of the monomer on the particle size and morphology. PVP microspheres were prepared in 0.23–0.74 μm size range with Monasil PCA and 0.71–1.98 μm size range with KF-6017, respectively. The resulting polymer particle of >90% yield was obtained. Particle size slightly increased with the amount of monomer in polymerization with Monasil PCA. In the case of KF-6017 as the surfactant, there was not an obvious variation in particle size with increasing monomer. Particle size of PVP decreased as surfactant concentration increased from 5.0 to 15.0 wt.% basis on concentration of monomer. The narrow particle size distribution (Dn = 0.23 μm and PSD = 1.06) was presented at the high concentration of Monasil PCA (15 wt.% on monomer concentration). As indicated by the reaction temperature and the addition of organic solvent, which affected solubility of monomer, polymer and surfactant in scCO2, particle size and particle size distribution of PVP varied. PVP particles with Monasil PCA strongly aggregated at 75 °C in contrast to KF-6017 which showed discrete particles at 65 and 70 °C, but particle size distribution was broad. Particle size was slightly reduced with a little amount of hexane, with an inverse relationship of adding hexane reduced the particle size. The amount of the relative residual surfactants on surface of the polymer after extracting with supercritical fluid process (SFE) was measured by using SEM/EDS and EPMA analysis to map out the distribution of silicon element qualitatively. The original polymer particle before the extraction using CO2 had the high level of silicon element, but the average level of silicon element became low after CO2 extraction.  相似文献   

4.
《Diamond and Related Materials》2007,16(4-7):1121-1125
We have found that several precious metal-loaded MgO catalysts are active in the formation of carbon nanotube (CNT) by the chemical vapor deposition (CVD) of methane. The catalysts were prepared with nine metals (Fe, Co, Ni, Ru, Rh, Pd, Os, Ir, and Pt) by impregnation onto a high surface area MgO. CNT synthesis was carried out in the temperature range from 600 °C to 1000 °C after reduction with H2 at 800 °C.The amount of carbon deposited and crystallinity in the produced CNT on nine metals showed interesting tendencies: (i) The amount of carbon formed increased in the following transition series metals: first < second < third row transition elements, and (ii) the index of crystallinity IG/ID in Raman-bands of the CNTs decreased in the following order: 8 > 9 > 10 in the Periodic Table. Group 8 and 9 metals produced tube type fibers composed of the graphite layers arranged parallel to the fiber axis. On the other hand, carbon nanofibers (CNFs) grown on group 10 metals had herringbone type graphene sheets.  相似文献   

5.
A mix-valenced nickel oxide, NiOx, was prepared from nickel nitrate aqueous solution through a precipitation with sodium hydroxide and an oxidation by sodium hypochlorite. Further, pure nickel oxide was obtained from the NiOx by calcination at 300, 400 and 500 °C (labeled as C300, C400 and C500, respectively). They were characterized by thermogravimetry (TG), X-ray diffraction (XRD), nitrogen adsorption at −196 °C and temperature-programmed reduction (TPR). Their catalytic activities towards the degradation of phenol were further studied under continuous bubbling of air through the liquid phase. Also, the effects of pH, temperature and kinds of nickel oxide on the efficiency of the microwave-enhance catalytic degradation (MECD) of phenol have been investigated. The results indicated that the relative activity affected significantly with the oxidation state of nickel, surface area and surface acidity of nickel oxide, i.e., NiOx (>+2 and SBET = 201 m2 g−1)  C300 (+2 and SBET = 104 m2 g−1) > C400 (+2 and SBET = 52 m2 g−1) > C500 (+2 and SBET = 27 m2 g−1). The introduction of microwave irradiation could greatly shorten the time of phenol degradation.  相似文献   

6.
The solid solutions of BiFeO3–BaTiO3 have been prepared via solid state with a view to obtaining magnetoelectric properties, i.e. ferroelectric and magnetic activity in the same range of temperatures. Optimum calcination and sintering strategy for obtaining pure perovskite phase, dense ceramics (>97% relative density) and homogeneous microstructures have been determined. The sample of composition 0.7BiFeO3–0.3BaTiO3 reported in the present work is pseudo-cubic at room temperature. The permittivity is ɛr ≈ 150 at the room temperature and shows a broad ferro-para phase transition at around 175 °C where ɛr  1600. This diffuse maximum of the permittivity, similar to that in relaxors, is due to the chemical inhomogeneity in both A and B sites of the perovskite unit cell ABO3. Higher losses, tan δ > 1, appear above 200 °C and other different conduction mechanisms start to be active particularly at temperatures higher than 400 °C, when the ceramic becomes conductive. The magnetic properties show a succession of transitions from weak ferro/ferrimagnetism-to-antiferromagnetism and antiferromagnetism-to-paramagnetism at TN1  10 K and TN2  265 K. Below TN2 the ceramic 0.7BiFeO3–0.3BaTiO3 can present magnetoelectric coupling, due to the fact that is simultaneously ferroelectric and antiferromagnetic.  相似文献   

7.
The TiO2 ceramics were prepared by a solid-state reaction in the temperature range of 920–1100 °C for 2 h and 5 h using TiO2 nano-particles (Degussa-P25 TiO2) as the starting materials. The sinterability and microwave properties of the TiO2 ceramics as a function of the sintering temperature were studied. It was demonstrated that the rutile phase TiO2 ceramics with good compactness could be readily synthesized from the Degussa-P25 TiO2 powder in the temperature range of 920–1100 °C without the addition of any glasses. Moreover, the TiO2 ceramics sintered at 1100 °C/2 h and 920 °C/5 h demonstrated excellent microwave dielectric properties, such as permittivity (Ɛr) value >100, Q × f  > 23,000 GHz and τf  200 ppm/°C.  相似文献   

8.
Carbon-supported Pt–Sn/C bimetallic nanoparticle electrocatalysts were prepared by the simple reduction of the metal precursors using ethylene glycol. The catalysts heat-treated under argon atmosphere to improve alloying of platinum with tin. As-prepared Pt–Sn bimetallic nanoparticles exhibit a single-phase fcc structure of Pt and heat-treatment leading to fcc Pt75Sn25 phase and hexagonal alloy structure of the Pt50Sn50 phase. Transmission electron microscopy image of the as-prepared Pt–Sn/C catalyst reveals a mean particle diameter of ca. 5.8 nm with a relatively narrow size distribution and the particle size increased to ca. 20 nm when heat-treated at 500 °C due to agglomeration. The electrocatalytic activity of oxygen reduction assessed using rotating ring disk electrode technique (hydrodynamic voltammetry) indicated the order of electrocatalytic activity to be: Pt–Sn/C (as-prepared) > Pt–Sn/C (250 °C) > Pt–Sn/C (500 °C) > Pt–Sn/C (600 °C) > Pt–Sn/C (800 °C). Kinetic analysis reveals that the oxygen reduction reaction on Pt–Sn/C catalysts follows a four-electron process leading to water. Moreover, the Pt–Sn/C catalyst exhibited much higher methanol tolerance during the oxygen reduction reaction than the Pt/C catalyst, assessing that the present Pt–Sn/C bimetallic catalyst may function as a methanol-tolerant cathode catalyst in a direct methanol fuel cell.  相似文献   

9.
Combustive oxidation of volatile organic compounds (VOCs), such as propyl alcohol, toluene and cyclohexane, were studied. The combustion was catalyzed by nanoparticles of La1−xSrxCoO3 (x = 0, 0.2) perovskites prepared by a co-precipitation method. The results showed high activities of the perovskite catalysts. Compared to LaCoO3, in particular, La0.8Sr0.2CoO3 was much higher in catalytic ability. The total oxidation of VOCs followed the increasing order: cyclohexane < toluene < propyl alcohol. The T99% of cyclohexane was 40 °C lower than that of toluene, which appeared to be determined by the bond strengths of the weakest C–H and C–C bonds. The 100-h stability experiments showed that La1−xSrxCoO3 (x = 0, 0.2) perovskite was highly stable.  相似文献   

10.
Two 1D coordination polymeric isomers with the same chemical formula [Cu(2,2′-bipy)(Htda)]n (2,2′-bipy = 2,2′-bipyridine, Htda = 1,2,3-triazole-4,5-dicarboxylate) were solvothermally prepared and structurally characterized. X-ray analysis revealed that one isomer has a 1D staircase chain structure while the other possesses a 1D spiral chain structure. The electrochemical behaviors of both isomers in solution were also investigated by cyclic voltammetry.  相似文献   

11.
The mechanism of the hydrogenation of aliphatic C4–C6 dinitriles (succinonitrile, glutaronitrile, and adiponitrile) over Raney-type Ni catalysts was investigated in a fed-batch autoclave at 350 K and 5.0 MPa. The results are interpreted based on the strength of adsorption, the interaction with the solvent and intramolecular interactions. The kinetics of the hydrogenation of aliphatic dinitriles is highly dependent on the hydrocarbon chain length. Short dinitriles like succinonitrile adsorb stronger on the catalyst surface than longer dinitriles like adiponitrile. The yield of the intermediate aminonitriles decreases with increasing hydrocarbon chain length, due to the enhanced competitiveness of dinitrile and aminonitrile for the same active sites. It is proposed that the reactivity of dinitriles and aminonitriles is caused by difference in solvent interaction. It is remarkable that the stronger the adsorption, the higher the reactivity. The reactivity decreases in the order: succinonitrile > glutaronitrile > adiponitrile. In contrast, the reaction rate of the aminonitriles is fairly independent of the hydrocarbon chain length. The formation of undesired secondary amines in the form of cyclic compounds is under kinetic control and increases with decreasing chain length. It is found that adiponitrile can be very selectively hydrogenated to primary diamines. Promoting Raney-type Ni catalysts with traces of Mo, Cr or Fe enhances the performance in the hydrogenation of dinitriles. This gives the best opportunity for process improvement towards the desired primary amines.  相似文献   

12.
We present here the single-source-precursor synthesis of Fe3Si and Fe5Si3-containing SiOC ceramic nanocomposites and investigation of their magnetic properties. The materials were prepared upon chemical modification of a hydroxy- and ethoxy-substituted polymethylsilsesquioxane with iron (III) acetylacetonate (Fe(acac)3) in different amounts (5, 15, 30 and 50 wt%), followed by cross-linking at 180 °C and pyrolysis in argon at temperatures ranging from 1000 °C to 1500 °C. The polymer-to-ceramic transformation of the iron-modified polysilsesquioxane and the evolution at high temperatures of the synthesized SiFeOC-based nanocomposite were studied by means of thermogravimetric analysis (TGA) coupled with evolved gas analysis (EGA) as well as X-ray diffraction (XRD). Upon pyrolysis at 1100 °C, the non-modified polysilsesquioxane converts into an amorphous SiOC ceramic; whereas the iron-modified precursors lead to Fe3Si/SiOC nanocomposites. Annealing of Fe3Si/SiOC at temperatures exceeding 1300 °C induced the crystallization of Fe5Si3 and β-SiC. The crystallization of the different iron-containing phases at different temperatures is considered to be a consequence of the in situ generation of a Fe–C–Si alloy within the materials during pyrolysis. Depending on the Fe and Si content in the alloy, either Fe3Si and graphitic carbon (at 1000–1200 °C) or Fe5Si3 and β-SiC (at T > 1300 °C) crystallize. All SiFeOC-based ceramic samples were found to exhibit soft magnetic properties. Magnetization versus applied field measurements of the samples show a saturation magnetization up to 26.0 emu/g, depending on the Fe content within the SiFeOC-based samples as well as on the crystalline iron silicide phases formed during pyrolysis.  相似文献   

13.
A visible-light-active N-containing TiO2 photocatalysts were prepared from crude amorphous titanium dioxide by heating amorphous TiO2 in gaseous NH3 atmosphere. The calcination temperatures ranged from 200 to 1000 °C, respectively. UV–vis/DR spectra indicated that the N-doped catalysts prepared at temperatures <400 °C absorbed only UV light (Eg = 3.3 eV), whereas samples prepared at temperatures ≥400 °C absorbed both, UV (Eg = 3.10–3.31 eV) and vis (Eg = 2.54–2.66 eV) light. The chemical structure of the modified photocatalysts was investigated using FT-IR/DRS spectroscopy. All the spectra exhibited bands indicating nitrogen presence in the catalysts structure. The photocatalytic activity of the investigated catalysts was determined on a basis of a decomposition rate of nonionic surfactant (polyoxyethylenenonylphenol ether, Rokafenol N9). The most photoactive catalysts were those calcinated at 300, 500 and 600 °C. For the catalysts heated at temperatures of 500 and 600 °C Rokafenol N9 removal was equal to 61 and 60%, whereas TOC removal amounted to 40 and 35%, respectively. In case of the catalyst calcinated at 300 °C surfactant was degraded by 54% and TOC was removed by 35%. The phase composition of the most active photocatalysts was as follows: (a) catalyst calcinated at 300 °C—49.1% of amorphous TiO2, 47.4% of anatase and 3.5% of rutile; (b) catalyst calcinated at 500 °C—7.1% of amorphous TiO2, 89.4% of anatase and 3.5% of rutile; (c) catalyst calcinated at 600 °C—94.2% of anatase and 5.8% of rutile.  相似文献   

14.
With the aim of multifunctionalized all-optical photorefractive polymers bearing a non-conjugated main chain (NCMC), four novel norbornene-containing monomers were synthesized by introduction of different alkyl spacers between the functional chromophore and norbornene group. The polymers were prepared via ring-opening metathesis polymerization (ROMP) and characterized by IR, UV–Vis, GPC, DSC, TG, etc. Their number average molecular weights (Mn) are higher than 1.0 × 104, and their glass transition temperatures are in the range from 75 °C to 153 °C depending on the volume of functional group and spacer length. The polymers have good thermal stability up to 280 °C and excellent solubility in THF. Optically transparent films were fabricated with N-ethyl-carbazole (ECZ) as a plasticizer and C60 as a charge generator. Their photorefractive performance was evaluated at 633 nm by two-beam-coupling (TBC) experiment at a zero electrical field. TBC signals were observed from these films without prepoling, indicating that the synthesized NCMC polymers bear an all-optical photorefractive property.  相似文献   

15.
The gelling temperature of Poloxamer 407/Poloxamer 188 (16%/10%) mixture solution increased with increasing amount of surfactants. Among surfactants tested, the order of effectiveness in increasing the gelling temperature was sodium dodecyl sulfate (SDS) > tween 20 > cetyltrimethylammonium chloride (CTAC). The gelling temperature increased with the increasing concentration of cyclodextrin (CD). The order of effectiveness in increasing the gelling temperature was gamma CD > beta CD > alpha CD. The gelling temperature increased with increasing concentration of alcohols (ethanol and propylene glycol (PG)). On the contrary, dimethyl sulfoxide (DMSO) decreased the gelling temperature of P407/P188 (16%/10%) mixture solution. On the other hand, the gelling temperature decreased upon addition of NaCl but it increased by addition of MgCl2. With P407/P188-loaded monoolein cubic phase, the degrees of release in 40 hr were about 52% at 25 °C and 37.5% at 47 °C. The suppressed release at 47 °C is possibly because P407/P188 in the water channel of cubic phase is in gel state at the temperature.  相似文献   

16.
Barium strontium titanate, (BaxSr1?x)TiO3 (BST) thin films have been prepared on alumina substrate by sol–gel technique. The X-ray patterns analysis indicated that the thin films are perovskite and polycrystalline structure. The interdigital electrode with 140 nm thickness Au/Ti was fabricated on the film with the finger length of 80 μm, width of 10 μm and gaps of 5 μm. The temperature dependence of dielectric constant of the BST thin films in the range from ?50 °C to 50 °C was measured at 1 MHz. The dielectric properties of the BST thin films were measured by HP 8510C vector network analyzer from 50 MHz to 20 GHz.  相似文献   

17.
Thermally stable and insoluble silver salts of 12-molybdophosphoric acid with varying amount of Ag cations were prepared. XRD results indicated the presence of single phase of AgxH3−xPMo12O40 (0 < x < 3). FTIR and Raman results indicated that Ag was incorporated in the secondary structure of Keggin ion. The catalytic conversion of ethanol increased in the order of H3PMo > Ag2PMo > Ag1PMo > Ag3PMo. Pure H3PMo is highly selective to dimethyl ether formation; in contrast AgxH3−xPMo12O40 catalysts offered better ethylene selectivity. The pyridine adsorption studies revealed that increase of Ag incorporating led to increase the Lewis acid sites.  相似文献   

18.
A series of diphenylanthrazolines were synthesized by Friedländer condensation of 2,5-dibenzoyl-1,4-phenylenediamine and acetyl-functionalized compounds in the presence of polyphosphoric acid as catalyst, in yields ranging from 61% to 88%. The diphenylanthrazolines are thermally robust with high decomposition temperatures (>380.0 °C) and high melt transitions (317–462 °C). All of them show the lowest energy absorption bands (λmaxAbs: 394–433 nm) from the π–π1 transitions by virtue of their large molar extinction coefficients (?  104 M?1 cm?1), revealing low optical band gaps (2.59–2.80 eV). The compounds emit blue fluorescence with λmaxEm ranging from 430 to 466 nm in dilute toluene solution.  相似文献   

19.
The effect of macromolecular architecture on the aggregative behavior of AB block copolymers with different complex structures in selective solvents was studied by Dissipative Particle Dynamics. We focus on two types of diblock copolymers, (I) asymmetric linear diblocks ByAx and (II) miktoarm stars (By)n(Ax)m, where A block is solvophilic and B block solvophobic. Note that y and x are the block lengths of A and B blocks; n and m denote arm numbers of A and B blocks in the star. For type I linear copolymer with a given ratio of y/x, the aggregation number <p> varies with the total length (x + y) for y/x > 1 but is essentially independent of the total length for y/x ? 1. For type II star copolymer with m · x = 24 and n · y = 24, the aggregation number varies with the branch number m at a given number of solvophobic blocks n. There exists a minimum aggregation number at m  4 so that <p> declines first and then grows with increasing the branch number. Moreover, <p> increases as the polymer concentration is increased. Our simulation results indicate that at a given chemical composition, the micelle properties such as aggregation number and micellar morphology may vary with the macromolecular architecture.  相似文献   

20.
Lead zirconium titanate [Pb(ZrxTi1?x)O3 or PZT] thin films were prepared by the thermal annealing of multilayer films composed of binary oxide layers of PbO, ZrO2 and TiO2. The binary oxides were deposited by metal organic chemical vapor deposition. An interdiffusion reaction for perovskite PZT thin films was initiated at approximately 550 °C and nearly completed at 750 °C for 1 h under O2 annealing atmosphere. The composition of Pb/Zr/Ti in perovskite PZT could be controlled by the thickness ratio of PbO/ZrO2/TiO2 where the contribution of each binary oxide at the same thickness was 1:0.55:0.94. The electrical properties of PZT (Zr/Ti = 40/60, 300 nm) prepared on a Pt-coated substrate included a dielectric constant ?r of 475, a coercive field Ec of 320 kV/cm, and remnant polarization Pr of 11 μC/cm2 at an applied voltage of 18 V.  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