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.
以精萘、浓硫酸和甲醛为原料,依次经过磺化、水解、缩合三个反应合成了萘磺酸甲醛缩合物(NSF),并获得了各反应的优化条件。结果表明:当2-萘磺酸(2-NSA)产率最大时,磺化反应的优化条件为n(萘)∶n(浓硫酸)=1∶1.15、反应温度160℃、反应时间3 h;水解反应要使1-萘磺酸(1-NSA)的残留量最低优化条件为反应温度115℃、反应时间60 min、n(萘)∶n(水)=1∶2.3,此时1-NSA水解的转化率最高;缩合反应的优化条件为n(萘)∶n(甲醛)=1∶1、酸度30%,反应温度105℃,反应时间2 h,在此条件下2-NSA全部聚合生成NSF。产物NSF的红外光谱(FTIR)和高效液相色谱(HPLC)分析表明,不同条件下获得的NSF都含有丰富的芳环、亚甲基和磺酸基结构;HPLC分析表明,NSF在磺化度、分子量大小和分子的主体结构(线性结构/枝状结构)方面存在差异,这种差异导致NSF在作为水煤浆分散剂时,对降低浆体黏度和提高浆体稳定性方面具有不同的性能;优化条件下得到的NSF在降低浆体黏度方面性能优于市售的萘磺酸甲醛缩合物系分散剂(NX-1),但二者对保持浆体稳定性的能力相当。  相似文献   

2.
《分离科学与技术》2012,47(8):2103-2116
Abstract

Rotating bead mill and baffled bioreactors have earlier been shown to provide excellent mass transfer and bioremediation rates for naphthalene particulates. In this study, the mass transfer rates of naphthalene and methylnaphthalenes from NAPL into water in both the bead mill and baffled bioreactors are reported. The values of KLa ranged between 1.0 h?1 and 42 h?1, similar to values observed with suspended PAH particulates, increasing with bead loadings up to 50% by volume, bead size up to 5.0 mm, rotation rate up to 50 RPM, oil loading up to 72 mL (7.2% volume fraction) and naphthalene loading up to 1000 mg/L (based on the water phase). Baffled bioreactors provided similar volumetric mass transfer coefficients as bead mill bioreactors, but without the loss of working volume due to the presence of solid beads.  相似文献   

3.
A novel and sensitive electrochemical sensor based on the cone conformation of the supramolecule 25, 27-(3-thiopropoxy)-p-tert-butyl calix[4]arene has been developed for quantitative determination of polycyclic aromatic hydrocarbons (PAHs). The method works effectively by immobilizing calix[4]arenes on Fe3O4 magnetic nanoparticles. CdSe quantum dots were used as electrochemical labels. CdSe quantum dots (QDs) modified PAHs in competition with the sample PAHs were intercalated into calix[4]arenes supramolecules via a host–guest interaction through individual bowl-shaped calix[4]arenes. The stripping analysis of the cadmium dissolved from CdSe nanoparticles provided a sensitive method for the detection of PAHs in the samples. The signal decrease of the QDs was proportional to the increase in the concentration of the PAHs. Under optimal conditions, among the five PAHs, the square wave voltammetry (SWV) response of QDs decreased linearly for anthracene and naphthalene in the range of 2.1 × 10?7–1.4 × 10?5 and 1.5 × 10?6–2.5 × 10?5 M, respectively. The calculated detection limits (3δ) were 20.1 ng mL?1 for anthracene and 105.5 ng mL?1 for naphthalene.  相似文献   

4.
Summary A one-pot method is described for the synthesis of naphthalene oligomers polyalkylated with 1-octadecene and 1-hexadecene. The synthesis involves two steps: 1) oligomerization of naphthalene using an AlCl3/CuCl2 catalyst/oxidant system, and 2) polyalkylation with a 1-olefin. Characterization studies include1H- and13C-NMR spectroscopy, GPC and testing methods for lubricants. The degree of alkylation, variations in the alkyl groups, and the position of the substituent on the naphthalene ring were established by13C NMR spectroscopy. The purified products possess excellent properties for being suitable as a lubricant base stock, lubricant additives and functional fluids.  相似文献   

5.
This work deals with the dealumination of zeolite beta under different conditions to optimize its catalytic performance. The dealumination was carried out either using oxalic or tartaric acid solution of different pH values or steaming up to 500 °C. The dealuminated zeolite samples were characterized using X-ray diffraction, N2-physisorption, ICP, 27Al NMR and NH3-TPD. Their catalytic activities were studied for acylation reaction of 2-methoxy naphthalene or naphthalene with acetic anhydride and esterification of benzyl alcohol with hexanoic acid. Among the dealuminated samples examined, those treated with oxalic acid or tartaric acid at pH 2 showed high activity. The high activity can be correlated with the relative increase in the amount of strong acid sites and with the enhanced accessibility of the reactants to the active sites, which are caused by removal of extra-framework Al species on the acid treatments.  相似文献   

6.
Activated carbon fibers (ACF) are prepared from phenol–formaldehyde resin fibers through chemical activation and physical activation methods. The chemical activation process consisted of KOH, whereas the physical activation was performed by activation in CO2. The characteristics of the electrochemical supercapacitors with carbon fibers without activation (CF), carbon fibers activated by CO2 (ACF-CO2), and carbon fibers activated by KOH (ACF-KOH) have been compared. The activated carbon fibers from phenol–formaldehyde resins present a broader potential range in aqueous electrolytes than activated carbon and other carbon fibers. Activation does not produce any important change in the shape of starting fibers. However, activation leads to surface roughness and larger surface areas as well as an adapted pore size distribution. The higher surface areas of fibers treated by KOH exhibited higher specific capacitances (214 and 116 F g−1 in aqueous and organic electrolytes, respectively) and good rate capability. Results of this study suggest that the activated carbon fiber prepared by chemical activation is a suitable electrode material for high performance electrochemical supercapacitors.  相似文献   

7.
A sulfonic acid functionalized titanium dioxide quasi-superparamagnetic nanocatalyst Fe3O4@SiO2@TiO2-OSO3H with average size of 61 nm and semispherical shape with surface area about 97 m2 g?1 with saturation magnetization 17.7 emu g?1 and the coercivity 9.84 Oe was successfully synthesized. The structure and morphology of the nanocatalyst was characterized by Fourier transform infrared spectroscopy (FT-IR), energy-dispersive X-ray spectroscopy, X-ray diffraction pattern, transmission electron microscopy, field-emission scanning electron microscopy, vibrating sample magnetometer and Brunauer–Emmett–Teller surface area analysis. The catalytic usage of the nanocatalyst was exemplified in synthesis of 2,3-dihydroquinazolin-4(1H)-one and spiroquinazolin-4(3H)-one derivatives in deep eutectic solvents (DESs) based on choline chloride and urea. We suggest that the synergistic effects in catalytic activities of titanium dioxide, organic acid and the CO2 capture property of DES are the main reasons for the improvement of catalytic activity. The synthesized spiroquinazolinones and dihydroquinazolinones derivatives were characterized by FT-IR, 1H and 13C nuclear magnetic resonance spectroscopy. The magnetic nanocatalyst exhibit high catalytic activity and can be simply separated from reaction media by an external magnet in a few seconds and could be reused for six cycles without significant loos in activity, which indicates the good immobilization of sulfonic acid on the magnetic titanium dioxide support. Furthermore, the solvent which has been used in this work can be readily isolated and reused for several times.  相似文献   

8.
Isonicotinate self-assembled monolayers (SAM) were prepared on alumina surfaces (A) using isonicotinic acid (iNA). These functionalized layers (iNA-A) were used for the seeded growth of copper films (Cu-iNA-A) by hydrazine hydrate-initiated electroless deposition. The films were characterized by scanning electron microscopy (SEM), electron-dispersive X-ray spectroscopy, atomic force microscopy, X-ray photoelectron spectroscopy, X-ray diffraction, and advancing contact angle measurements. The films are Cu0 but with surface oxidation, and show a faceted morphology, which is more textured (R q = 460 ± 90 nm) compared to the SAM (R q = 2.8 ± 0.5 nm). In contrast, growth of copper films by SnCl2/PdCl2 catalyzed electroless deposition, using formaldehyde (CH2O) as the reducing agent, shows a nodular morphology on top of a relatively smooth surface. No copper films are observed in the absence of the isonicotinate SAM. The binding of Cu2+ to the iNA is proposed to facilitate reduction to Cu0 and create the seed for subsequent growth. The films show good adhesion to the functionalized surface.  相似文献   

9.
ABSTRACT

Extraction of 241Am and 154Eu by 50% tributyl phosphate (TBP) in C13 fraction from nitric acid solutions containing 6 mol L?1 nitrate-ion of ammonium, aluminum, and iron salts has been investigated. Rig trials with transplutonium elements (TPE) and fission rare earth (RE) recovery from the evaporated raffinate of reprocessed WWER-1000 spent fuel were carried out in hot cells using mixer-settlers. The HLW feed acidity was adjusted by dissolving metallic iron up to 1 mol L?1 Fe(NO3)3. TPE and RE were recovered by 99.98% with DF from 137Cs > 7500. The combined extraction-partitioning cycle of a Modified Purex-process is proposed using 50% TBP as the single solvent.  相似文献   

10.
Single crystals of gossypol with three molecules of acetic acid (gossypol triacetic acid) were grown from solutions of gossypol acetic acid and acetone. The crystals were unstable in air but could be stabilized for X-ray diffraction analysis by coating the crystal surfaces with a thin layer of mineral oil. The gossypol triacetic acid complex (C30H30O8·3C2H4O2) forms an orthorhombic crystal system with P212121 (Z=4) symmetry. Unit cell dimensions were a=9.0208(7) Å, b=17.4884(10) Å, and c=24.358(2), Å yielding a volume of 3842.7(5) Å3 and a density of 1.2077(2) g/cm3. As with all previously reported crystals of gossypol, the gossypol molecules were of the aldehyde tautomer, and the two planar naphthalene rings were approximately perpendicular. Acetic acid molecules were found to lie in channels within the gossypol matrix. Individual crystals contained only one gossypol enantiomer, but both enantiomers crystallized from solution. Although the crystal habit could not be used to distinguish between the gossypol enantiomers, a fragment of the crystal could be derivatized and analyzed by high-performance liquid chromatography for this purpose. The ability to grow large, nonracemic crystals leads to a simple procedure for separating small quantities of the individual gossypol enantiomers.  相似文献   

11.
Product flexibility is key to meeting fluctuating chemicals demands in the future. In this contribution, the methanol to hydrocarbons (MTH) reaction was investigated over two Ge-containing H-ITQ-13 samples, one with needle-like (H-ITQ-13(N), with (Si+Ge)/Al) = 42) and another with plate-like (H-ITQ-13(P), with (Si+Ge)/Al > 100) morphology. The samples were characterised using XRD, BET, SEM/EDS and FTIR spectroscopy, and their MTH performance was compared with the performance of H-ZSM-5 and H-ZSM-22. Similar specific surface areas (413 and 455 m2 g?1 for H-ITQ-13(N) and (P), respectively) and similar acid strength (Δν ~ ?327(?310) cm?1) was observed for the two H-ITQ-13 samples. Testing of H-ITQ-13(N) at weight hourly space velocity (WHSV) = 2–8 h?1 at 350–450 °C revealed that C5+ alkenes were the main products (35–45 % selectivity at 400 °C), followed by propene and butene. A low but significant selectivity for aromatic products was observed (6–8 % selectivity at 400 °C). Product selectivity was found to be independent of deactivation. The methanol conversion capacity of H-ITQ-13(N) was 120–150 g methanol g?1 catalyst at 400 °C. Testing H-ITQ-13 at high (30 atm) and ambient pressure, respectively, at 350 °C showed that a high pressure led to enhanced C5+ selectivity, but close to a tenfold decrease in methanol conversion capacity. H-ITQ-13(P) was tested at 400 °C and 2 h?1. It gave lower conversion than H-ITQ-13(N). Furthermore, when compared at the same conversion level, H-ITQ-13(P) gave higher C5+ alkene selectivity, lower aromatics selectivity, and a higher propene to ethene ratio than H-ITQ-13(N). The H-ITQ-13 samples yielded a product spectrum intermediate of H-ZSM-22 and H-ZSM-5. The effluent product cut-off of H-ITQ-13 was similar to that of H-ZSM-5 with tetramethylbenzene as the largest significant product, while H-ZSM-22 produced mainly linear and branched alkenes. The lifetime of H-ITQ-13(N) was clearly enhanced compared to H-ZSM-22, but inferior to H-ZSM-5.  相似文献   

12.
13C-NMR spectroscopy has been applied to the analysis of formaldehyde containing resins. Samples have been prepared from phenols, urea and melamine with formaldehyde. The chemical shifts of the carbon atoms in the resins were measured relative to TMS. All the samples could be dissolved in DMSO-d6 which facilitated the comparison of signals in different resins. The spectra were interpreted with the aid of spectra of reference compounds and simple calculations based on additivity increments. Each type of resin gives a specific 13C-NMR spectrum in which lines can be assigned that give information on the structure of the resin in relation to the type of condensation or catalyst used. Moreover, the different ways in which formaldehyde is incorporated in the resins as methylene carbon can be unambiguously determined from the region between 20 and 100 ppm. This work shows that 13C-NMR spectroscopy is a powerful tool for the analysis of formaldehyde containing resins.  相似文献   

13.
《分离科学与技术》2012,47(13):2013-2019
Separation of U(VI) from Th(IV) and rare earth elements (REEs) present in monazite leach solution (nitric acid medium) has been studied using tris(2-ethylhexyl) phosphate (TEHP) and tri-n-butyl phosphate (TBP) dissolved in n-paraffin as solvents under varying experimental conditions such as nitric acid, extractant and metal ion concentrations etc. There is an increase in distribution ratio of U(VI) (D U ) with increase in aqueous phase acidity up to 5 M HNO3 beyond which a decrease is observed. Typically for 1 × 10?3 M U(VI), the DU values increase from 8 (0.5 M HNO3) to 80 (5 M HNO3) for 1.1 M TEHP, and from 2 (0.5 M HNO3) to 43 (5 M HNO3) for 1.1 M TBP in n-paraffin. The separation factors of U(VI) (β: DU/DM) over metal ions (M) such as Th(IV) and Y(III) (chosen as a representative of heavy REEs) are better for TEHP than TBP at all nitric acid concentrations. Batch solvent extraction data have been used to construct the McCabe-Thiele diagrams for the recovery of U(VI) employing TEHP as the extractant. A process flow sheet has been proposed with 0.2 M TEHP in n-paraffin as solvent for the recovery of U(VI) from simulated monazite leach solution in HNO3 medium.  相似文献   

14.
The reported acetate‐derived labelling of the fungal naphthalene γ‐pyrone fonsecin, two streptomycete dodecaketide αpyrones TW93f and TW93g, and the streptomycete phenanthraquinones piloquinone, murayaquinone and haloquinone appear to be exceptions to the generalisation that fungi and streptomycetes produce fused‐ring aromatic polyketides by different modes of cyclisation. A review of their 1) originally assigned formulae, 2) [13C2]acetate‐derived labelling patterns, and 3) modes of cyclisation leads to the recognition of feasible alternative chemical structures or biosynthetic pathways, which are in accord with the originally proposed classification system.  相似文献   

15.
Synthesis of 6,13-di (p-hydroxyphenyl) pentacene (PP-H) and 6,13-di (p-hydroxy napthyl) pentacene (PN-H) using 6, 13-pentacenequinone (PENTA) is reported. The PP-H and PN-H obtained were identified by 1H NMR, and infrared spectra. Substitution with hydroxy phenyl group or hydroxy Napthyl group at the C-6 and C-13 positions of pentacene leads to phenomenal enhancement in solubility and photooxidative stability. XRD results showed that the pattern of PP-H and PN-H was different from the patterns of PENTA. UV-Visible spectra showed that the λmax of PP-H and PN-H in CHCl3. The fluorescence spectra showed that PP-H (449 nm) and PN-H (396 nm) emitted purple when exited by UV radiation while only emitted red (PP-H 597 nm and PN-H 616 nm) when exited with visible light.  相似文献   

16.
The characteristics of urea-formaldehyde (UF) adhesives condensed by catalysis with four different acids, namely formic (HCOOH), hydrochloric (HCl), phosphoric (H3PO4), and sulfuric (H2SO4) acids, under alkaline–acidic–alkaline conditions at a molar ratio F/U = 1.12 were studied. The thermal curing properties of UF adhesives catalyzed with acid were characterized by differential scanning calorimetry at 10 °C/min heating rate. The resin structure examined by 13C-NMR spectroscopy showed that the resin catalyzed with HCl had a lower proportion of methylol groups, resulting in a lower level of formaldehyde emission. It was interesting to note that HCOOH resulted in the best overall mechanical properties of the medium-density fiberboard (MDF) panels. The HCl catalyst resulted in the poorest performance, providing the lowest internal bond strength, modulus of elasticity, and thickness swelling, with the exception of the free formaldehyde content. The resin catalyzed with H2SO4 had the highest free formaldehyde and the highest formaldehyde emission. H2SO4 and H3PO4 resulted in MDF mechanical properties relatively lower than for HCOOH. © 2018 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2019 , 136, 47256.  相似文献   

17.
Mesoporous nickel oxide (NiO) nanoparticles were synthesized by the thermal decomposition reaction of Ni(NO3)2·9H2O using oxalic acid dihydrate as the mesoporous template reagent. The pore structure of nanocrystals could be controlled by the precursor to oxalic acid dihydrate molar ratio, thermal decomposition temperature and thermal decomposition time. The structural characteristic and textural properties of resultant nickel oxide nanocrytals were characterized by X-ray diffraction (XRD), transmission electron microscopy (TEM), scanning electron microscopy (SEM), N2 adsorption–desorption isotherm and temperature programmed reduction. The results showed that the most excellently mesoporous nickel oxide particles (m-Ni-1-4) with developed wormlike pores were prepared under the conditions of the mixed equimolar precursor and oxalic acid and calcined for 4 h at 400 °C. The specific surface area and pore volume of m-Ni-1-4 are 236 m2 g?1 and 0.42 cm3 g?1, respectively. Over m-Ni-1-4 at space velocity = 20,000 mL g?1 h?1, the conversions of toluene and formaldehyde achieved 90 % at 242 and 160 °C, respectively. It is concluded that the reactant thermal decomposition with oxalic acid assist is a key step to improve the mesoporous quality of the nickel oxide materials, the developed mesoporous architecture, high surface area, low temperature reducibility and coexistence of multiple oxidation state nickel species for the excellent catalytic performance of m-Ni-1-4.  相似文献   

18.
Ceria catalysts have been prepared by a nanocasting procedure using SBA-15, MCM-48 and KIT-6 silica-based templates, and investigated for the total oxidation of naphthalene. In all cases cubic fluorite CeO2 was prepared, and the structure of the template was replicated when SBA-15 and MCM-48 were used. The KIT-6 template was not replicated by the nanocasting synthesis, but in all cases mesoporous CeO2 was obtained with high surface areas (91–190 m2 g?1). All of the catalysts demonstrated high activity for naphthalene oxidation to CO2, and the most active was the catalyst prepared from the KIT-6 template. The high activity was attributed to the small crystallite size of the CeO2, combined with high surface area and the highly accessible catalyst surface.  相似文献   

19.
A new technique of synthesis of mesoporous silica with protonated amino groups avoiding microwave treatment of mesophase was developed using a template method. The block-copolymer P123 was used as a template and sodium meta-silicate with 3-aminopropyltriethoxysilane as precursors. After the removal of template from mesophase with boiling ethanol, the obtained sample displayed highly ordered hexagonal structure with attractive textural parameters: SBET = 460 m2 g?1, Vtotal = 0.79 cm3g?1 and d = 7.1 nm. FTIR and 13C CP/MAS NMR spectroscopy revealed the presence of alkyl ammonium groups (0.7 mmol g?1) that were able to attach anions of molybdophosphoric acid to the surface of the synthesized mesoporous material. The resulting anion-ion exchange phase can find applications in many areas (adsorption, catalysis, etc.).  相似文献   

20.
Aerogel sulfated ZrO2–SiO2 mixed oxide solid acid catalyst was prepared by sol–gel method followed by supercritical drying (SCD) in n-propanol solvent, which resulted into higher surface area (170 m2/g), pore volume (0.31 cm3/g) and pore diameter (7.2 nm) having ordered mesoporous structure as well as more number of Brönsted and Lewis acid sites available on larger surface area. The catalyst exhibited 91 % yield of methyl stearate at 60 °C in 7 h, which increased from 71 to 91 % with an increase in the Zr to Si ratio from 1:2 to 2:1 due to increase in acid site concentration. The reaction followed pseudo-first order kinetics under the optimized reaction conditions with a reaction rate of 1.15 mmol h?1, rate constant of 2.7 × 10?1 h?1 and turn over frequency of 9.68 h?1. The catalyst displayed higher activity (91 %) compared to ion exchange resins (44–68 %), Nafion (58 %), acid clay (61 %) and pure sulfated zirconia (78 %), and was slightly lower as compared to H2SO4 (97 %). The study clearly reveals the improved structural, textural and acidic properties of ZrO2–SiO2 mixed oxide aerogel prepared via SCD technique.  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