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 312 毫秒
1.
The effects of a series of short chain alcohols, 1‐butanol (C4OH), 1‐pentanol (C5OH), and 1‐hexanol (C6OH), on the styrene (ST) emulsion polymerization mechanisms and kinetics were investigated. The CMC of the ST emulsions stabilized by sodium dodecyl sulfate (SDS) first decreases rapidly and then levels off when the CiOH (i = 4, 5, or 6) concentration ([CiOH]) increases from 0 to 72 mM. Furthermore, at constant [CiOH], the CMC data in decreasing order is CMC (C4OH) > CMC (C5OH) > CMC (C6OH). The effects of CiOH (i = 4, 5, and 6) on the ST emulsion polymerization stabilized by 6 mM SDS are significant. This is attributed to the reduction in CMC by CiOH, the different oil–water interfacial properties, the different concentrations of monomer within latex particles, and the different effectiveness of SDS/CiOH in stabilizing latex particles. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 100: 4406–4411, 2006  相似文献   

2.
The five independent stiffness constants, C11, C33, C44, C66, and C13, and the axial and transverse thermal expansivity of unidirectional gel-spun polyethylene fiber reinforced composites have been measured as functions of fiber volume fraction Vf. The axial extensional modulus C33 and axial Poisson's ratio v13 follow the rule of mixtures, while the axial shear modulus C44, transverse shear modulus C66, and transverse plane-strain bulk modulus Ct ( = C11C66) obey the Halpin-Tsai equation. Extrapolation to Vf = 1 gives the five stiffness constants of gel-spun polyethylene fiber. The tensile property of the fiber is highly anisotropic, with the axial Young's modulus about 40 times higher than the transverse Young's modulus. In contrast, the axial shear modulus exceeds the transverse shear modulus by only 5%. A similar treatment of the thermal expansivity data in terms of the Schapery equations gives an axial thermal expansivity of −1.25 × 10−5 K−1 and a transverse thermal expansivity of 11.7 × 10−5 K−1 for the fiber.  相似文献   

3.
Reaction of the complexes (SM,RC)‐[(η5‐C5Me5)M{(R)‐Prophos}(H2O)](SbF6)2 (M=Rh, Ir) with α,β‐unsaturated aldehydes diastereoselectively gave complexes (SM,RC)‐[(η5‐C5Me5)M{(R)‐Prophos}(enal)](SbF6)2 which have been fully characterized, including an X‐ray molecular structure determination of the complex (SRh,RC)‐[(η5‐C5Me5)Rh{(R)‐Prophos}(trans‐2‐methyl‐2‐pentenal)](SbF6)2. These enal complexes efficiently catalyze the enantioselective 1,3‐dipolar cycloaddition of the nitrones N‐benzylideneaniline N‐oxide and 3,4‐dihydroisoquinoline N‐oxide to the corresponding enals. Reactions occur with excellent regioselectivity, perfect endo selectivity and with enantiomeric excesses up to 94 %. The absolute configuration of the adduct 5‐methyl‐2,3‐diphenylisoxazolidine‐4‐carboxaldehyde was determined through its (R)‐(−)‐α‐methylbenzylamine derivative.  相似文献   

4.
A method for estimating absolute concentrations of C2H5 and H radicals in hydrocarbon diffusion flames is proposed and substantiated. Concentration profiles of C2H5 and H on the flame axis are obtained. In the method proposed, the concentration of C2H5 is determined from the equality of two quantities — the rate of loss of n-butane by diffusion and the rate of its formation by recombination of two C2H5 radicals. The concentration of H radicals is determined from the relation between the ratio C2H5/H and experimental profiles of C2H4, C2H6, and O2. __________ Translated from Fizika Goreniya i Vzryva, Vol. 43, No. 6, pp. 13–20, November–December, 2007.  相似文献   

5.
In addition to the ω-5 olefinic acids found in otherGrevillea species, about 10% of the acyl groups ofG. decora seed oil contain a hydroxy group and an ω-5 double bond. The chainlengths of these acids are from C22 to C30, with the largest concentration at the C26 and C28 chainlengths. The hydroxy group is located on odd carbons from carbon-5 through carbon-13. These acids previously were unknown in nature. The most abundant of these are 7-hydroxy-cis-17-docosenoic, 7-hydroxy-cis-19-tetracosenoic, 9-hydroxy-cis-19-tetracosenoic, 9-hydroxy-cis-21-hexacosenoic, 11-hydroxy-cis-21-hexacosenoic, and 13-hydroxy-cis-23-octacosenoic acids. The oil also contains the largest known concentration of the unoxygenated C26 and C28 ω-5 monoenes.  相似文献   

6.
P. E. Kolattukudy 《Lipids》1970,5(4):398-402
Surface lipid of pea leaves (Pisum sativum var. Frosty) was analyzed with column, thin layer and gas liquid chromatography in conjunction with mass spectrometry and infrared spectroscopy. It contained 42%n-hentriacontane and 7.3%n-hentriacontan-16-ol. About 5% was wax esters, C40–C50 consisting of primarily C26 and C28 alcohols and C16–C22 acids. Almost 5% was aldehydes, mainly C26 and C28. Primary alcohols, chiefly C26 and C28, made up 20% of the surface lipid.  相似文献   

7.
A series of alkyl sulphobetaine Gemini surfactants Cn‐GSBS (n = 8, 10, 12, 14, 16) was synthesized, using aliphatic amine, cyanuric chloride, ethylenediamine, N,N′‐dimethyl‐1,3‐propyldiamine and sodium 2‐chloroethane sulfonate as main raw materials. The chemical structures were confirmed by FT‐IR, 1H NMR and elemental analysis. The Krafft points differ markedly with different carbon chain length, for C8‐GSBS, C10‐GSBS and C12‐GSBS are considered to be below 0 °C and C14‐GSBS, C16‐GSBS are higher than 0 °C but lower than room temperature. Surface‐active properties were studied by surface tension and electrical conductivity. Critical micelle concentrations were much lower than dodecyl sulphobetaine (BS‐12) and decreased with increasing length of the carbon chain from 8 to 16, and can reach a minimum as low as 5 × 10?5 mol L?1 for C16‐GSBS. Effects of carbon chain length and concentration of Cn‐GSBS on crude oil emulsion stability were also investigated and discussed.  相似文献   

8.
9.
The interaction and synergism of some polyoxyethylenated fatty alcohol ether (POE) nonionic surfactants (C12E2, C12E3, C10E5, C10E7, where Cx indicates number of carbon atoms in the chain and Ey indicates number of oxyethylene glycol ethers) with trioxyethylenated dodecyl sulfonate (C12E3S) in mixed monolayer formation at the surface and in mixed micelle formation in aqueous solutions were studied at 25 and 40°C by calculating interaction parameters (βα, βM) from surface tension-concentration data by use of Rosen's equations based on the nonideal solution theory. All the systems investigated adapt reasonably well to the nonideal model, with negative values of βσ and βM (where M means micelle and σ refers to the air-liquid interface) indicating a favorable interaction between the mixed surfactants. Either at a monolayer or in a mixed micelle, the attractive interaction becomes stronger when the alkyl chain in the POE surfactant is longer, i.e., when the POE becomes more hydrophobic. The interaction increases in the order C10E7<C10E5<C12E3, C12E2. For the two C10E n (n= 5,7)/C12E3S systems, as temperature increases from 25 to 40°C, the interaction increases in a mixed micelle, but it decreases in a mixed monolayer. Synergism in mixed micelle formation exists for C12E3S/C10E n mixtures when X1 M , the mole fraction of POE in a mixed micelle, is ≈0.4–0.8, whereas synergism does not occur in the systems of C12E3S/C12E m due to the large difference between CMC1 and CMC2, i.e., large |In(C 1 M /C 2 M )| value (where CMC=critical micelle concentration). The degree of synergism in mixed micelle formation is temperature independent and is 0.23, 0.18, and close to zero for C10E5/C12E3S, C10E7/C12E3S, and C12E m (m=2,3)/C12E3S systems, respectively. Synergism in surface tension reduction effectiveness occurs in C12E3S/C12E2 and C12E3S/C12E3 systems. The mole fractions of POE in the solution phase are 0.302 and 0.333 for the two mixtures at the point of maximum synergism.  相似文献   

10.
Two novel late transition metals complexes with bidentate O?N chelate ligand, Mt(benzocyclohexan‐ketonaphthylimino)2 {Mt(bchkni)2: bchkni ?C10H8(O)C[N(naphthyl)CH3]; Mt ? Ni, Pd}, were synthesized. In the presence of B(C6F5)3, both complexes exhibited high activity toward the homo‐polymerization of norbornene (NB) (as high as 2.7 × 105 gpolymer/molNi·h for Ni(bchkni)2/B(C6F5)3 and 2.3 × 105 gpolymer/molPd·h for Pd(bchkni)2/B(C6F5)3, respectively). Additionally, both catalytic systems showed high activity toward the copolymerization of NB with 1‐octene under various polymerization conditions and produced the addition‐type copolymer with relatively high molecular weights (0.1–1.4 × 105g/mol) as well as narrow molecular weight distribution. The 1‐octene content in the copolymers can be controlled up to 8.9–14.0% for Ni(bchkni)2/B(C6F5)3 and 8.8–14.6% for Pd(bchkni)2/B(C6F5)3 catalytic system by varying comonomer feed ratios from 10 to 70 mol %. The reactivity ratios of two monomers were determined to be r1‐octene = 0.052, rNB = 8.45 for Ni(bchkni)/B(C6F5)3 system, and r1‐octene = 0.025, rNB = 7.17 for Pd(bchkni)/B(C6F5)3 system by the Kelen‐TÜdÕs method. The achieved NB/1‐octene copolymers were confirmed to be noncrystalline and exhibited good thermal stability (Td > 400°C, Tg = 244.1–272.2°C) and showed good solubility in common organic solvents. © 2012 Wiley Periodicals, Inc. J. Appl. Polym. Sci., 2013  相似文献   

11.
Using precise experimental data on the specific heat at constant pressure C p of the normal alkanes C7-C11 over a wide parameter range, the C p values are determined on the liquid and vapor branches of the boundary curve from the normal boiling point to 0.98T c. In the coordinates C p-P s-T s, orthogonal projections of the reduced phase diagram of these hydrocarbons are constructed and analyzed. Existing generalized expressions for C p of the liquid and gas phases of the normal alkanes C5-C11 on the boundary curve as a function of temperature are evaluated, and new expressions for C p are proposed.__________Translated from Teoreticheskie Osnovy Khimicheskoi Tekhnologii, Vol. 39, No. 4, 2005, pp. 476–480.Original Russian Text Copyright © 2005 by Kuznetsov, Gorbachev.  相似文献   

12.
Thermal decomposition of hexanitrohexaazaisowurtzitane (HNIW) was investigated through tuneable vacuum ultraviolet photoionization with molecular‐beam sampling mass spectrometry (MBMS). According to photoionization efficiency (PIE) spectroscopic results, the initial decomposition products of HNIW were identified including HCN, CO, NO, HNCO, N2O, CO2 (a little), NO2, C2H2N2, C3H3N3, C4H3N3, C3H4N4, C5H4N4, C5H5N5 and C6H6N6. The possible ionization energies of C2H2N2, C4H3N3, C3H4N4 and C6H6N6 were analyzed on basis of the PIE spectra. The data were compared with those of thermogravimetry‐mass spectrometry (TG‐MS) and thermogravimetry‐Fourier transform‐infrared spectroscopy (TG‐FT‐IR). The kinetic parameters for the formation of HNCO, HCN and CO2 were calculated from the current curves of species by TG‐FT‐IR spectroscopy, typically the apparent activation energy (Ea) and prefactor (A) for HNCO were Ea=161.3 ± 2.5 kJ mol−1 and A=38.9 ± 0.6 s−1 with an optimal mechanism function f(α)=(1−α). Global thermal decomposition reaction and Arrhenius equation of HNIW were suggested at the end.  相似文献   

13.
The volatile components of the Dufour's gland secretion of female halictid bees have been examined in 18 Nearctic species belonging toAgapostemon, Augochlora, Augochlorella, Augochloropsis, Dialictus, Evylaeus, Halictus, andLasioglossum. Nine saturated and unsaturated macrocyclic lactones ranging from C18 to C26 have been identified. Four of these compounds, the saturated C26 and the unsaturated C20, C22, and C24 lactones, are new natural products reported for halictine bees. A series of eight esters containing branched C5-alkenols and fatty acids has been identified in several species. The cell linings and pollen ball inAugochlora pura pura contain the same major lactones as the Dufour's gland. A discussion of the significance of the Dufour's gland secretion for apoid systematics and its function in the Halictidae is presented.  相似文献   

14.
The Ziegler catalyst, a combination of TiCl4 and Al(C2H5)2Cl, is a very poor catalyst for propylene polymerization. However, the addition of Mg(C4H9)2 to the Ziegler catalyst at a [Mg]:[Al] molar ratio < 0.5 leads to the formation of a potent catalyst system for homopolymerization of propylene. Polymers produced with the TiCl4–Al(C2H5)2Cl/Mg(C4H9)2 catalyst are mostly atactic and amorphous, their minor partially crystalline isotactic components constitute merely 5–10% of the combined polymer material. Principal advantages of the TiCl4–Al(C2H5)2Cl/Mg(C4H9)2 catalyst for the manufacture of atactic PP include high activity, low cost, and the ease of use: the catalyst is prepared in situ from three commercially available components. © 2019 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2019 , 136, 47692.  相似文献   

15.
A new analytical fundamental equation of state (EOS) is presented for fluids. The equation is explicit in the effective molecular potentials and allows calculation of all thermodynamic properties over the whole fluid surface (gas, liquid, supercritical and gas–liquid phase transition). Outside the critical area (± 0.05Tc), it is valid in a vast range of temperature and pressure (0.8Tc to 7.5Tc and up to 120Pc,). The EOS is applicable for a variety of refrigerants such as C3H2F4 (HFO-1234ze (E)), hydrofluorocarbons (HFCs) including C3F8 (R218), C3H2F6 (R236ea), C3H2F6 (R236fa), C3H3F5 (R245ca), C3HF7, C4F8 (RC318), C4F10, C5F12, natural refrigerants including NH3, CO2, hydrocarbons, monatomic gases and some other fluids. Calculations of second derivatives properties of fluids are sensitive tests of EOS behavior. Therefore, estimation of the thermodynamic properties including Joule-Thomson coefficient, μJT, and speed of sound, w, has been considered.  相似文献   

16.
N-(α-Carboxyalkyl)acrylamide telomer-type surfactants (xC n−1 AmAc where n is alkyl chain length=6, 8, 10, 12; and x is degree of polymerization=3.3–13.1) were synthesized by the telomerization of monomer (C n−1 AmAc) in the presence of the corresponding alkanethiol as a chain transfer agent and then investigated for their surface-active properties. xC n−1 AmAc telomers lowered the surface tension of aqueous solutions that were at pH 9–10. The critical micelle concentrations (CMC) of the telomers were lower than those of the monomers with the same alkyl chain length, and the CMC values shifted to lower concentrations with both increasing alkyl chain length and polymerization degree. xC9AmAc with x=3.3–6.3 gave the highest efficiencies in lowering the surface tension. The cross-sectional molecular areas per molecule of xC n−1 AmAc telomers were smaller than the values estimated on the assumption that they are assemblies of C n−1 AmAc monomer units. The foaming abilities and the foam stabilities were both in the orders of xC7AmAc>xC9AmAc>xC5AmAc>xC11AmAc. Mixtures of aqueous solutions of xC n−1 AmAc telomers and toluene formed oil-in-water emulsions. The emulsion-stabilizing abilities were in the orders of xC7AmAc>xC5AmAc>xC9AmAc=xC11AmAc. The addition of Ca2+ to the mixed solutions of telomers and toluene resulted in formation of water-in-oil type emulsions. Thus, the surface-active properties of the telomers were influenced significantly by the alkyl chain length and the polymerization degree of the telomers. In addition, these properties could be correlated with the hydrophilic-lipophilic balance (HLB); the highest surface activities were observed by using xC n−1 AmAc with HLB of 14–18.  相似文献   

17.
Anhydrous butterfat treated with copper and stored under vacuum at 38C in the dark for 9 weeks yielded the C5–8 n-alkanes, acetic acid and the C4,5,7,9 alkan-2-ones. The same butterfat stored under oxygen for 1 week yielded the C1,2,3,5-7 n-alkanes, the C2,4–6,8 alk-l-enes, the C7,8 alk-l-ynes, the C2–6 n-alkanoic acids, ethanol, the C2–7 n-alkanals, 3-methylbutanal and 4-methylpentanal, the C1–4 n-alkyl formates, ethyl acetate and propionate,n-propyl acetate and propionate, and butanone. Both treatments yielded carbon dioxide. All samples including those not treated with copper yielded butane-2,3-dione. Visiting Scientist from Division of Dairy Research, C.S.I.R.O., Highett, Victoria, Australia.  相似文献   

18.
Dialkyl hydroxypropyl sulfobetaine (HSB) surfactants, C16GA-(PO)5-(EO)3-HSB and C24GA-(PO)10-(EO)10-HSB, were synthesized from Guerbet alcohols (GA) polyoxypropylene–polyoxyethylene (PO-EO) ethers and their behaviors in surfactant-polymer (SP) flooding of high temperature and high salinity reservoirs were examined and compared with their anionic hydroxypropyl sulfonate (HS) counterparts, C16GA-(PO)5-(EO)3-HS and C24GA-(PO)10-(EO)10-HS. The PO-EO chain embedded improves their aqueous solubility, and the sulfobetaines show better salt resistance than sulfonates. For a reservoir condition of total salinity 19,640 mg L−1 and 60–80°C, C16GA-(PO)5-(EO)3-HSB alone can reduce crude oil/connate water interfacial tension (IFT) to ultralow at 0.25–5 mM, which can be further widened to 0.1–5 mM by mixing with dodecylhexyl (C12+6) glyceryl ether hydroxypropyl sulfobetaine (C12+6GE-HSB), a slightly hydrophobic surfactant. C24GA-(PO)10-(EO)10-HSB is more hydrophobic for the specified reservoir condition, however, by mixing with hexadecyl dimethyl hydroxypropyl sulfobetaine (C16HSB), a hydrophilic surfactant, ultralow IFT can also be achieved at a total concentration of 0.25–5 mM. The anionic counterparts can also reduce IFT to ultralow by mixing with C12+6GE-HSB and C16HSB, respectively. Moreover, the optimum binary mixture, C16GA-(PO)5-(EO)3-HSB/C12+6GE-HSB at a molar fraction ratio of 0.6/0.4, can keep the negatively charged solid surface water-wet (θw = 12–23°) in a wide concentration range, and can still achieve ultralow IFT after stored at 90°C for 90 days (initially 5 mM), which overall are favor of improving oil displacement efficiency at high temperature and high salinity reservoir conditions.  相似文献   

19.
The reaction of [RhCl(PiPr3)2] ( 1 ) with 1,4-C6H4(C≡CH)2 at 0°C leads almost quantitatively to the formation of the bis(alkyne) complex [(PiPr3)2ClRh-(HC≡C6H4-C≡CH)RhCl(PiPr3)2] (2). At elevated temperatures (THF, 60°C) it rearranges to give the isomeric bis(vinylidene) complex [(PiPr3)2ClRh-(=C=CH-C6H4-CH=C=)RhCl(PiPr3)2] (3). A one-pot synthesis of 3 is also described. Treatment of either 2 or 3 with pyridine affords the bis(alkynyl)dihydrido compound [(PiPr3)2(py)Cl(H)Rh(-C≡C-C6H4-C≡C-)Rh(H)Cl(py)(PiPr3)2] ( 4 ) in which both metal centers are octahedrally coordinated. Whereas the reaction of 2 with NaC5H5 produces the complex CsH5(PiPr3)Rh(HC≡C-C6H4-C≡CH)Rh(PiPr3)C5H5 ( 7 ), the bis(vinyl-idene) isomer C5H5(PiPr3)Rh(=C=CH-C6H4-CH=C=)Rh(PiPr3)C5Hs ( 8 ) is obtained from 4 and NaC5H5. Electrophiles preferably attack the Rh=C bonds of 8 and thus on protonation with CF3CO2H the bis(vinyl) complex C5H5(PiPr3)(CF3CO2)-Rh(Z,Z-CH=CH-C6H4-CH=CH)Rh(O2CCF3)(PiPr3)C5Hs ( Z-9 ) is formed. In acetone solution, it rearranges to give the E isomer. Reaction of 8 with sulfur affords the bis(thioketene) complex C5H5(PiPr3)Rh(≡2-C,S; η2-C,S-S=C=CH-C6H4-CH=C=S)-Rh(PiPr3)C5H5 ( 12 ), for which only one diastereomer is observed. All attempts to prepare mononuclear rhodium compounds containing the diyne HC≡C-C6H4-G≡CH or the isomeric vinylidene: C=CH-C6H4-G≡CH as ligand failed.  相似文献   

20.
Anhydrous butterfat was irradiated at 6 megarads in a special glass reaction flask, and the headspace and total condensate samples were examined by wide-range (−180C to 125C) gas chromatography combined with mass spectrometry. The nature and amounts of volatile compounds were not greatly influenced by whether the butterfat was irradiated under oxygen or under vacuum, nor (apart fromn-alkanoic acids) by storage for 1 and 9 weeks. Carbon dioxide was produced in greatest amount. Of the remaining compounds, aliphatic hydrocarbons predominated both in number and amount. Aliphatic oxygenated compounds including carbonyl compounds were isolated in relatively small amounts. The following compounds were identified positively: C1–13 n-alkanes; C4–9 2-methylalkanes; C2–14 alk-l-enes; C2–9 alk-l-ynes; C2–5 n-alkanoic acids; C1–5 n-alkan-l-ols and propan-2-ol; C2–8 n-alkanals and 2-methylbut-2-enal; C3–5,7 alkan-2-ones; C1–4 n-alkyl formates, vinyl and isoamyl formate, methyl acetate and methyln-hexanoate; and carbon dioxide. Visiting Scientist from Division of Dairy Research, C.S.I.R.O., Highett, Victoria, Australia.  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