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 15 毫秒
1.
Sodium α-sulfonated, fatty acid polyethylene glycol monoesters [C m H2m+1CH(SO3Na)COO(C2H4O) n H] and diesters [C m H2m+1CH(SO3Na)COO(C2H4O) n COCH(SO3Na)C m H2m+1], wherem=10–16 andn=1–35, were prepared by esterification of α-sulfonated, fatty acids with polyethylene glycols, followed by neutralization with NaOH. Crude products were purified by reversed-phase column chromatography on an octadecyl-modified silica gel. Characteristic solution behavior of these α-sulfonated fatty acid esters was, examined, and the following features were observed. All monoesters prepared in this work had Krafft points below 0°C and also possessed good calcium stabilities. Critical micelle concentrations of the monoesters increased monotonously, as a rule, with an increase in the number of oxyethylene units. These results suggest that the polyethylene glycol residue of the monoester behaves as a hydrophile. On the other hand, diesters possessed high water solubility, low foamability, and critical micelle concentrations that were lower by a factor of ten compared to those of the monoesters.  相似文献   

2.
Two-phase Hydroformylation of Buta-1,3-diene and Hydrocarbon Mixtures Containing Buta-1,3-diene The two-phase hydroformylation of buta-1,3-diene with (HRh(CO)[P(m-C6H4SO3Na)3]3 the Kuntz catalyst system with excess P(m-C6H4SO3Na)3) gives high yields of C5-monoaldehydes. Main product in this mixture is the reactive trans- and cis-pent-3-enal. In consecutive reactions the pent-3-enal is partially hydrogenated to n-pentanal, but also-favoured by the protolytic milieu of the two-phase reaction- aldol condensated to 2-propenylheptadienal. The hydrogenation product of the propenylheptadienal, 2-propylheptanol-1, is a good plasticizer alcohol with a wanted low vapour pressure. Especially promising is the two-phase hydroformylation of the unrefined C4-fraction of the naphtha pyrolysis: after a more than 95 per cent conversion of the buta-1,3-diene also more than 80 per cent of the n-but-1-ene in the C4-fraction is hydroformylated mainly to wanted n-pentanal. Less than 5–10% of the n-but-2-enes and the isobutene in the C4-fraction react under these conditions to oxo-products (2- and 3-methylbutanal). Acetylenic compounds in the C4-fraction are converted quantitatively into products.  相似文献   

3.
To better control the interfacial activities of selenium‐containing surfactants, a redox‐switchable anionic surfactant containing 2 selenium atoms, namely sodium 3‐((3‐(benzylselanyl)propyl)selanyl)propyl sulfate (SBSe2S), has been designed and synthesized. Upon oxidation by H2O2, the 2 divalent selenide groups in the hydrophobic tail of SBSe2S are converted into the corresponding selenoxides. The initial hydrophobic tail of SBSe2S gets separated into 3 segments by 2 hydrophilic selenoxide groups, destroying the bola‐type structure. The interfacial activities of SBSe2S in aqueous solution could thereby be switched off to a greater degree than in the case of its counterparts containing only a single selenium atom. After reduction with Na2SO3, the 2 selenoxide groups in SBSe2S‐Ox are restored to the initial selenide. Consequently, the interfacial activities of SBSe2S could be reversibly switched by alternate addition of H2O2 and Na2SO3, without obvious deterioration over 5 cycles.  相似文献   

4.
1-(4-Isobutylphenyl)ethanol (IBPE) was carbonylated to 2-(4-isobutylphenyl)propionic acid (ibuprofen) in an aqueous/organic two phase system using the water-soluble Pd(tppts)3 catalyst [tppts = P(C6H4-m-SO3Na)3] in the presence of p-CH3C6H4SO3H at 363 K, 15 MPa CO pressure and a palladium concentration of 150 ppm without addition of organic solvents. Under these conditions the conversion of IBPE was 83% and the selectivity to ibuprofen 82% with no decomposition of the Pd(tppts)3 catalyst. Both the activity and selectivity were strongly influenced by the tppts/Pd molar ratio and the nature of the added Brønsted acid. Maximum efficiency was observed for P/Pd = 10. Acids of weakly or non-coordinating anions, such as p-CH3C6H4SO3H, CF3COOH or HPF6 afforded carbonylation. No catalytic activity was observed in the presence of acids of strongly coordinating anions, such as HI. The water-soluble Pd/dppps catalyst [dppps = Ar2-nPhnP-(CH2)3-PPhnAr2-n; Ar = C6H4-m-SO3Na; n = nń = 0: 86% and n = 0, nń = 1: 14%] exhibited low catalytic activity and the major product obtained was the linear isomer of ibuprofen, 3-(4-isobutylphenyl)propionic acid (3-IPPA) with selectivities up to 78%. Replacement of tppts by a ligand containing less −SO3Na groups such as monosulphonated triphenylphosphine (tppms) gives rise to a dramatic drop in the catalytic activity and selectivity to ibuprofen. No catalytic activity was observed using palladium catalysts modified with 2-pyridyldiphenylphosphine (PyPPh2) and tris(2-pyridyl)phosphine (PPy3) which are both water soluble in their protonated form. A catalytic cycle is proposed to explain the observed results. ©1997 SCI  相似文献   

5.
The capabilities of ion bombardment and laser ablation coupled to mass spectrometry as independent techniques to investigate the surface thermooxidative stability of polystyrene, polybutadiene polymers, and styrene butadiene rubber (SBR) copolymers were investigated. Surface chemical modifications were detected according to the polymeric structure. The degradation products detected by static secondary ion mass spectrometry appeared at m/z 29, 43, and 55. Their compositions were related to the general formulae CnHmO+ with n = 1–3 and m = 1–3 for polybutadiene and styrene butadiene copolymers, whereas polystyrene was not affected by the aging treatment. The CnHmO+ ions result from butadiene unit degradation. The laser ablation ionization Fourier transform ion cyclotron resonance mass‐spectrometry results confirmed the detection of CnHmO+ ions. Finally, it may be considered that the surface thermooxidative process of SBR copolymers begins with butadiene unit degradation. The development of butadiene unit oxidation showed a dynamic oxidation phase, which coincided with a loss of unsaturation. The influence of the polymer conformation (blocked, branched, and random) on the surface oxidation for 30% styrene SBR compounds was also studied. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 87: 1910–1917, 2003  相似文献   

6.
High pyroelectric performance around human body temperature is essential for ultra-sensitive infrared detectors of medical systems. Herein, toward human health monitoring, composite ceramics (1-x)Pb0.99Nb0.02[(Zr0.57Sn0.43)0.94Ti0.06]0.98O3/xAl2O3 (x = 0, 0.1, and 0.2) were designed. A metastable ferroelectric (FE) phase was induced in the anti-FE matrix by the Al2O3 component-induced internal stress, and in turn FE-anti-FE phase boundary was constructed. The ceramics at x = 0.2 exhibit high pyroelectric coefficient with p = 10.9 × 10−4 C·m−2·K−1 and figures of merit with current responsivity Fi = 6.23 × 10−10 m·V−1, voltage responsivity Fv = 12.71 × 10−2 m2·C−1, and detectivity Fd = 7.03 × 10−5 Pa−1/2 around human body temperature. Moreover, the enhanced pyroelectric coefficients exist in a broad operation temperature range with a large full width at half maximums of 18.5°C and peak value of 29.2 × 10−4 C·m−2·K−1 at 48.2°C. The designed composite ceramic is a promising candidate for infrared thermal imaging technology of noncontact human health monitoring system.  相似文献   

7.
Metalation of methallyl phenyl selenide, allyl phenyl selenide, and methallyl phenyl sulfide followed by tri-n-butylstannylation gave exclusively the γ-stannylated sulfide or selenide (e.g., 1-phenylseleno-2-methyl-3-tri-n-butylstannyl-1-propene ( 5 )). Compound 5 reacts with a series of immonium electrophiles to give products of aminomethylation α- to selenium. Electrophiles have included N-(bromomethyl)phthalimide/ZnBr2, Eschenmoser salt (dimethylmethyleneammonium iodide), and Mannich reagents generated in situ from diethylamine, piperidine, isopropyl sarcosinate and benzylmethylamine. The selenide from the last of these amines (N, 3-dimethyl-N-benzyl-2-phenylseleno-3-butenamine, 10 ) has been subjected to further transformations as follows: (1) treatment of 10 with trimethylstannyllithium results in replacement of phenylseleno by stannyl; this allyltin ( 11 ) can then again be aminomethylated giving compound 12 or 13 ; (2) oxidation of 10 gives amino alcohol 14 by [2, 3] sigmatropic rearrangement of the allyl selenoxide; (3) photolysis of 10 results in 1,3-rearrangement of the phenylseleno group; oxidative rearrangement of this allyl selenide gives amino alcohol 16 . The phthalimidomethylation product ( 6 ) is converted to a precursor for the side chain of the cytokinin zeatin by oxidation and [2, 3] sigmatropic rearrangement.  相似文献   

8.
Poly(di(ω-alkylphenyl)stannane)s, [Sn(C n H2n Ph)2] m with n = 2–4, and a copolymer of di(3-propylphenyl)stannane and dibutylstannane of weight-average molar masses of 2–8 · 104 g/mol were synthesized by dehydropolymerization of stannanes of the composition H2SnR2 using Wilkinson’s catalyst [RhCl(PPh3)3]. At least two methylene groups were required as spacers between the phenyl group and the tin atom for polymerization to occur. The polystannanes were characterized by, among other techniques, 1H, 13C and 119Sn NMR spectroscopy, thermal analysis and X-ray diffraction. The polymers featured properties different from those of the corresponding poly(dialkylstannane)s. Specifically, the [Sn(C n H2n Ph)2] m family displayed glass transitions at remarkably low temperatures, down to ca. −50 °C, and a lower value for a copolymer (−68 °C). Polymers [Sn(CnH2nPh)2]m with n = 2 and 3 and a copolymer at room temperature were of a gel-like concistence, which enabled facile orientation with shear forces. Finally, the temperature-dependent electrical conductivity was determined for poly(di(3-propylphenyl)stannane), which followed the law of typical semiconductors, with an activation energy for conduction of 0.12 eV.  相似文献   

9.
Yo-ping Greg Wu  Ya-fen Lin  Chang-Tang Chang 《Fuel》2007,86(17-18):2810-2816
The goal of this study is to find out the exhaust emissions differences produced by different kinds of fatty acid methyl esters (FAME) derived from used cooking oils and animal fats, as well as the importance of the purification step in exhaust emissions production. A total of 120 L of waste vegetable oil and 30 L of waste frying oil were collected and converted into three batches of FAME. There were two batches of FAME produced from waste vegetable oil (B01 and B02), and one batch of FAME produced by mixing 2% of waste frying oil with waste vegetable oil (B03). The FAMEs used in this study had higher density, kinematic viscosity, and flash point, but a lower gross heating value, when compared to the premium diesel. The B01 engine produced higher CO formation and the diesel-fuelled engine produced higher CO than the B02 and B03 did for engine speeds higher than 1400 rpm. Most of the FAME fuels produced higher CO2 than the diesel fuel did. The FAME fuels emitted higher NOx and PM, but lower SO2, than the diesel fuel. CnH2n+2, diphenyl sulfone (C12H10O2S), and diethyl phthalate (C12H14O4) can be selected as the character index for the combustion of FAME.  相似文献   

10.
The self‐aggregation behavior of catanionic surface active ionic liquids (SAIL), 1‐alkyl‐3‐methylimidazolium alkyl sulfates, [C4mim][C16SO4] and [C6mim][C14SO4], in aqueous solutions was explored by several techniques. These results were analyzed together with the concerning data of [C8mim][C12SO4] reported previously. The feature of these SAIL is that the total carbon number of the hydrophobic chains is kept constant, while only the length of the two hydrocarbon chains varies. Because of the different intermolecular and intramolecular interactions, the hydrophobic interaction of asymmetrical SAIL is enhanced, which explains the minimal CMC value of [C4mim][C16SO4]. Thermal motion of hydrocarbon chains at the air–water interface leads to higher Amin and lower Γmax values with the increasing asymmetry of two hydrophobic chains. The conductivity measurements reveal that the micellization process is spontaneous and entropy‐driven in the studied temperature range. This work indicates that the symmetry of alkyl chains can influence the aggregation behavior of [Cnmim][CmSO4] in aqueous solutions.  相似文献   

11.
A series of Fe2O3-doped (1–5%) Pt/SO42 /ZrO2 were prepared by a co-precipitation method. The incorporation of small amounts of Fe2O3 into Pt/SO42 /ZrO2 results in an enhanced Brønsted acidity in the presence of H2, which makes the Fe2O3-doped catalysts much more active than the undoped one for n-heptane hydroisomerization. A maximal yield of C7 isomers appears at 3.5% Fe2O3.  相似文献   

12.
The crystal structures, pyroelectric properties, and thermal stability of [111]-oriented 0.5 mol% Mn-doped 0.36Pb(In1/2Nb1/2)O3-0.36Pb(Mg1/3Nb2/3)O3-0.28PbTiO3 (Mn-0.36PIN-0.36PMN-0.28PT) ternary single crystal were investigated. The temperature dependence of the Raman spectra and dielectric properties revealed that the crystal exhibited a rhombohedral (R) structure at room temperature, and ferroelectric R → tetragonal (T) and ferroelectric T to paraelectric cubic (C) phase transitions at 130 and 175°C respectively. The single crystal had a high remnant polarization of Pr = 38 μC cm–2 and coercive field of EC = 12 kV cm–1 at room temperature and a frequency of f = 100 Hz. The values of Pr and EC decreased with increasing temperature, exhibiting anomalies near their phase-transition temperatures, which coincided with changes in the Raman spectra and dielectric properties. Furthermore, at 25°C and f = 100 Hz, the single crystal had high pyroelectric coefficients of p = 8.7 × 10−4 C m−2 K−1, figures of merit for the current responsivity of Fi = 3.5 × 10−10 m V−1, the voltage responsivity of Fv = 0.08 m2 C−1, and the detectivity of Fd = 30.1 × 10−5 Pa−1/2. These values were weakly dependent on temperature below 120°C. In addition, the room-temperature pyroelectric coefficients of the ternary single crystal maintain over 83% of the original value at thermal annealing temperatures below 120°C. These outstanding pyroelectric properties, together with high thermal stability, indicate that [111]-oriented rhombohedral Mn-0.36PIN-0.36PMN-0.28PT ternary single crystal is a new potential candidate for infrared detection applications.  相似文献   

13.
《Catalysis communications》2002,3(11):533-539
SO2 adsorption, SO2 oxidation and oxidation of propane with oxygen in the absence and the respective presence of SO2 in the feed gas were studied over unsulfated and sulfated 1% Pt/γ-Al2O3.Results showed that the promoting effect of SO2 in the reaction flux on C3H8 oxidation over 1% Pt/γ-Al2O3 depends on the presulfating temperature. Catalytic activity measurements and FTIR absorption spectra showed that during propane oxidation, Pt/support interfacial adsorbed species were formed at temperatures 25–300 °C, inhibiting C3H8 oxidation. However, at higher temperatures these Pt/support interfacial adsorbed species were oxidized, leading to Pt/support interfacial sulfate species, which strongly promote propane oxidation.  相似文献   

14.
Production of hydrogen by splitting of water in the thermochemical sulfur-based cycles that employs the catalytic decomposition of sulfuric acid into SO2 and O2 is of considerable interest. However, all of the known catalytic systems studied to date that consist of metal particles on oxide substrates deactivate with time on stream. To develop an understanding of the factors that are responsible for catalyst activity, we investigate the fresh activity of several platinum group metals (PGM) catalysts, including Pd, Pt, Rh, Ir, and Ru supported on titania at 850 °C and perform an extensive theoretical study (density-functional-theory-based first-principles calculations and computer simulations) of the activity of the PGM nanoparticles of different size and shape positioned on TiO2 (rutile and anatase) and Al2O3 (γ- and η-alumina) surfaces. The activity and deactivation of the catalytic systems are defined by (i) the energy barrier for the detachment of O atoms from the SOn (n = 1, 2, 3) species, and (ii) the removal rate of the products of the sulfuric acid decomposition (atomic O, S, and the SOn species) from metal nanoparticles. We show that these two nanoscale features collectively result in the observed experimental behavior. The removal rate of the reaction products is always lower than the SOn decomposition rates. The relation between these two rates explains why the “softer” PGM nanoparticles (Pd and Pt) exhibit the highest initial catalytic activity.  相似文献   

15.
The total oxidation of C2H2, C3H6, C3H8, n-C4H8, n-C4H10 and i-C4H10 was studied in a monolithic flow reactor under temperature-programmed mode and highly diluted conditions. The non-catalytic combustion of the investigated hydrocarbons leads to a substantial formation of CO and traces of methane at intermediate temperatures. This drawback was suppressed upon the deposition of a thin layer of Co3O4 on the monolith and all investigated hydrocarbons tend to light off at 250–290 °C. The apparent activation energy was found to exhibit a linear correlation with the C–H bond dissociation energy, indicating that the C–H activation is still the rate-limiting step.  相似文献   

16.
Anionic sulfonate gemini surfactants 8‐s‐8(SO3)2 and 12‐s‐12(SO3)2 (s = 3, 6) were synthesized, and their micellization in aqueous solution at 25.0 °C and pH 9 was investigated. The results show that the critical micelle concentrations (CMC) of 8‐s‐8(SO3)2 are more than 2 orders of magnitude larger than those of 12‐s‐12(SO3)2, but the spacer length has a relatively small impact on the CMC. Moreover, the interactions of nsn(SO3)2 (n = 8, 12; s = 3, 6) with anionic polyacrylamide (PAM) at 25.0 °C and pH 9 were investigated using surface tensiometry, rheolgy, and scanning electron microscopy (SEM). The results indicate that the surface tension and rheological properties of PAM depend on the concentration of nsn(SO3)2. Below the critical aggregation concentration of C1, surface tension is sharply reduced, the surface tension of nsn(SO3)2/PAM is lower than nsn(SO3)2 alone, but viscosity is almost unchanged with increasing Cnsn(SO3)2. Above C1, surface tension reduces very slowly until the saturated concentration of C2 is reached. Above C2, surface tension rapidly reduces until CM is attained, suggesting free nsn(SO3)2 micelles begin to form. In the region of C1CM, the viscosity significantly increases. Above CM, surface tension is basically unchanged and these curves coincide with those of the single surfactant system. Moreover, the viscosity is almost constant. The SEM images indicate that fibrous aggregates are formed below C1, then transformed into multilayer fibrous aggregates above C1, and further into fiber‐braided‐structured and spider‐web‐structured aggregates above CM. The variation of viscosity is closely associated with the transformation of aggregates.  相似文献   

17.
With the aim of preparing mullite, reactions between aluminum sulfate and silica in appropriate proportions and molten sulfate media M2SO4 (M=Na and/or K) were performed at different temperatures. The powders obtained were characterized by XRD, FT-IR, SEM and TEM. The reactivity was the same in Na2SO4 and (K,Na)2SO4 media. The best results in terms of yield (98.3%) and weight of mullite produced (95%) were obtained in Na2SO4 at 950 °C. The mullite phase exhibits an acicular morphology (75×0.75 μm) and a specific surface area close to 20 m2/g. In K2SO4 medium, a potassium alumino silicate is formed as well as mullite.  相似文献   

18.
Two environmentally friendly succinic acid monofluoroalkyl sulfonate surfactants were synthesized from maleic anhydride and polyethylene glycol mono (1H,1H,7H‐dodecafluoroheptyl) ether, i.e. H(CF2)6CH2OCH2CH2OCOCH(SO3Na)CH2COOH (FEOS‐1) and H(CF2)6CH2(OCH2CH2)3OCOCH(SO3Na)CH2COOH (FEOS‐3). The obtained surfactants were characterized by FT‐IR, 1H NMR, 13C NMR and 19F NMR in detail. The synthesized fluorinated surfactants have a high thermal stability on the basis of thermogravimetric analysis. Their surface properties were examined and the results show that FEOS‐1 and FEOS‐3 surfactants can reduce the surface tension of water to 25.55 mN m?1 at 10.25 mmol L?1 and 21.63 mN m?1 at 8.33 mmol L?1, respectively; meanwhile, the introduction of oxyethylene groups enhances the hydrophilicity and micellar forming ability and the longer oxyethylene chains the better surface properties. The Krafft points (Kp) of FEOS‐1 and FEOS‐3 were both below 0 °C, which was lower than perfluoro‐n‐heptanesulfonic acid sodium salt (n‐C7F15SO3Na, Kp = 56.5 °C) at a similar length of fluorocarbon chains. Comparison studies on two surfactants above and the conventional fluorocarbon surfactants, perfluorooctanoate of ammonium (PFOA) show that the surfactants have comparable properties to PFOA, thus offering an environmentally friendly synthesizing alternatives to PFOA.  相似文献   

19.
The microporous hydrogels (Pn‐Cm gels) composed of poly(dimethylaminoethyl methacrylate) and carboxymethylchitosan were synthesized in situ radical polymerization by using nano γ‐Fe2O3 particles as pore‐agent. The microporous structure formed through eliminating the Fe2O3 particles was designed to achieve a faster response rate and better drug loading effect. Comparing to the neat gels, Pn‐Cm gels exhibit deteriorative mechanical properties with the increased pores, while the gels still keep the elastic network structure which could bear some degree of tensile and compression deformation. Meantime, Pn‐Cm gels show similar temperature and pH double responsiveness with same isoelectric point shrink as that of neat gels, the swelling ability decreases slightly, and the deswelling rate increases with the increase of pores. Moreover, the 5‐fluorouracil was used as a target drug to explore the potential of this gel applied as drug‐release system. For Pn‐Cm gels, the more pores and carboxymethyl chitosan inside the gels are beneficial to the drug loading, all gels show a burst release of drug, being followed by a slow and sustained release with different rate. Comprehensively, the Pn‐Cm gels exhibit a better sustained release effect in the simulated stomach condition (pH = 2.1), the related release mechanism could be interpreted by the superposition of Fickian diffusion. © 2017 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2017 , 134, 45326.  相似文献   

20.
Polydichlorophosphazene [NPCl2]n reacts with the diphenol 2,2′‐(HO)C6H4‐C6H4(OH) in THF in the presence of K2CO3 to give the polymers [NP(O2C12H8) · x(OC4H8)]n (1) that contain variable ammouts of polytetrahydrofuran (PTHF) with x ranging from 0.05 to 0.8. This PTHF content (x) depends on the method followed to prepare the THF solutions of [NPCl2]n used for the reactions with the biphenol and can be made negligibly small, forming these solutions in the presence of K2CO3. This reveals the presence in the [NPCl2]n of acidic species capable of catalyzing the ring opening polymerization of THF. Polyphosphazene [NP(O2C12H8)]n (2) was prepared completely free of PTHF using dioxane as solvent. A comparison of the thermal behavior and morphological data of the polymers 1, 2, PTHF (5), and mixtures of [NP(O2C12H8)]n + x(OC4H8)m (4) revealed that the products 1 are strongly interacting polymer blends and ruled out the possibility of block copolymers. © 2000 John Wiley & Sons, Inc. J Appl Polym Sci 77: 568–576, 2000  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