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 109 毫秒
1.
The properties of alkyl sulfate and alkyl sulfonate are similar except for their Krafft points. However, alkyl sulfate and alkyl sulfonate behave quite differently when they are mixed with cationic surfactants and show some totally unexpected results. In this work sodium alkyl sulfate (CnH2n+1SO4Na,CnSO4)–alkyl quaternary ammonium bromide [CnH2n+1N(CmH2m+1)3Br, CnN, m=1–4] mixtures and sodium alkyl sulfonate (CnH2n+1SO3Na, CnSO3)–CnN mixtures were studied. It was found that, in contrast to the single surfactants, CnSO3–CnN mixtures were much more soluble than CnSO4–CnN mixtures. Besides, the two kinds of catanionic surfactant mixtures were quite different in their phase behavior and aggregate properties. The results were interpreted in terms of the interactions between surfactant molecules, which were very different in the two kinds of mixed systems owing to the distinction between alkyl sulfate and alkyl sulfonate in the molecular charge distribution.  相似文献   

2.
3‐Alkyl/aryl‐3‐hydroxyquinoline‐2,4‐diones were reduced with NaBH4 to give cis‐3‐alkyl/aryl‐3,4‐dihydro‐3,4‐dihydroxyquinolin‐2(1H)‐ones. These compounds were subjected to pinacol rearrangement by treatment with concentrated H2SO4, resulting in 4‐alkyl/aryl‐3‐hydroxyquinolin‐2(1H)‐ones. When a benzyl (Bn) group was present in position 3 of the starting compound, its elimination occurred during the rearrangement, and the corresponding 3‐hydroxyquinolin‐2(1H)‐one was formed. The reaction mechanisms are discussed for all transformations. All compounds were characterized by IR, 1H‐ and 13C‐NMR spectroscopy, as well as mass spectrometry.  相似文献   

3.
A series of 3‐(3‐hydroxyphenyl)‐4‐alkyl‐3,4‐dihydrobenzo[e][1,3]oxazepine‐1,5‐dione compounds with general formula CnH2n+1CNO(CO)2C6H4(C6H4OH) in which n are even parity numbers from 2 to 18. The structure determinations on these compounds were performed by FT‐IR spectroscopy which indicated that the terminal alkyl chain attached to the oxazepine ring was fully extended. Conformational analysis in DMSO at ambient temperature was carried out for the first time via high resolution 1H NMR and 13C NMR spectroscopy.  相似文献   

4.
Dichloro­(4,4′‐dipentyl‐2,2′‐bipyridine‐κ2N,N′)platinum(II), [PtCl2(C20H28N2)], adopts a discrete π–π stacking structure, where the alkyl chains are located in a random manner. In contrast, dichloro­(4,4′‐diheptyl‐2,2′‐bipyridine‐κ2N,N′)platinum(II), [PtCl2(C24H36N2)], forms a layer structure comprised of alkyl chain layers and paired coordination sites, as observed for analogous complexes with longer alkyl chains.  相似文献   

5.
Ibis paper reports the properties of the novel tetra‐p‐nitro‐tetra‐O‐alkyl‐calix[4]arenes (alkyl= n‐C4H9, 1; n‐C8H17 2; n‐C12H25, 3; n‐C16H33, 4). X‐ray crystallographic analysis and 1H NMR revealed that they exist as pinched‐cone conformation in crystal or cone conformation in solution. EFISH experiments at 1064 nm in CHCl3, indicated that tetra‐p‐nitro‐tetra‐O‐butyl‐calix[4]arene (1) has higher hyperpolarizability β, values than the corresponding reference compound p‐nitro‐phenyl butyl ether, without red shift of the charge transfer band. Compounds 2, 3 and 4 with longer alkyl chains can form monolayer at the air/water.  相似文献   

6.
Poly(2,3-dialkylbutanediol-1,4 terephthalates) with the alkyl substituents CH3, C2H5, n-C3H7, iso-C3H7, n-C4H9, and n-C10H21, andn-C16H33 were synthesized from the corresponding 2,3-dialkylbutanediols-1,4 and dimethyl terephthalate or terephthaloyl chloride. The substituents of the butanediol-1,4 portion of the polyterephthalates influence the 13C NMR chemical shifts of the carbon atoms near the branching site, the glass transition (Tg), and the crystallizability. Small alkyl substituents do not change the Tg of the polymers, whereas bulky substituents such as the isopropyl group increase the Tg and long normal alkyl groups as substituents decrease the Tg of the polymers. Crystallinity in these polyterephthalates was found only with CH3 and C16H33 as the 2,3-dialkyl substituents in the butanediol-1,4 portion of the polyester. This crystallinity of polyterephthalate of 2,3-di-C16H33 substituted butanediol-1,4 could be assigned to side-chain crystallization of the paraffinic groups.  相似文献   

7.
A convenient and general method for the regiospecific synthesis of three novel series of 1‐(2‐thenoyl)‐, 1‐(2‐furoyl)‐ and 1‐(isonicotinoyl)‐3‐alkyl(aryl)‐5‐hydroxy‐5‐trifluoromethyl‐4,5‐dihydro‐1H‐pyrazoles, in good yields (53 – 91 %), from the cyclocondensation reactions of 1,1,1‐trifluoro‐4‐alkoxy‐4‐alkyl(aryl)‐but‐3‐en‐2‐ones, where alkyl = H and Me; aryl = ‐C6H5, 4‐CH3C6H4, 4‐CH3OC6H4, 4‐FC6H4, 4‐ClC6H4, 4‐BrC6H4, 4‐NO2CgH4 with 2‐thiophenecarboxylic hydrazide, furoic hydrazide and isonicotinic acid hydrazide, respectively, is reported. Subsequently dehydration reaction of phenyl substituted 2‐pyrazolines with P2O5 furnished the corresponding 1H‐pyrazoles as mixture of regioisomers and in low yields (35 – 36 %).  相似文献   

8.
Fourier-transform ion-cyclotron-resonance (FTICR) mass spectrometry has been used to uncover the mechanisms by which FeO+ dehydrates heptan-4-one ( 5a ) and nonan-5-one ( 6a ) in the gas phase. The study of isotopomeric ketones provides evidence that H2O loss is not due to a 1,1-elimination, thus ruling out the intermediacy of high-valent iron-carbene species. Rather, H2O is generated in a formal 1,2-elimination involving the ω/ω ? 1 positions of the alkyl chain (‘remote C? H bond activation’). In the consecutive alkene/H2O elimination, the olefins (ethylene from 5a and propene from 6a ) originate from the terminal part of one alkyl chain, and the H-atom is transferred to the FeO+ moiety in the course of this process, builds up together with an H-atom from the ω/ω-1 position of the other alkyl chain the H2O molecule. In either case, the O-atom of H2O is provided by the FeO+ species.  相似文献   

9.
Using specific deuterium labelling the mechanisms of the olefin elimination reactions leading to formation of [C6H7]+ in the H2 and CH4 chemical ionizatin mass spectra of ethylbenzene and n-propylbenzene (and to [C2H5C6H6]+ in the CH4 chemical ionization mass spectra) have been investigated. The results show that the reaction does not occur by specific migration of H from the β position of the alkyl group to the benzene ring. For ethylbenzene 23–29% of the migrating H originates from the α-position, while for n-propylbenzene H migration from all propyl positions is observed in the approximate ratio, position 1:position 2:position 3=0.30:0.22:0.48. It is proposed that the results can be explained on the basis of competing H migration from each alkyl position involving cyclic transition states of different ring sizes, rather than by H randomization within the alkyl chain.  相似文献   

10.
Polymerization of vinylcyclohexane (VCHA) with TiCl3–aluminum alkyl catalysts was investigated. The polymerization rate of VCHA was low due to the branch at the position adjacent to the reacting double bond. The effects of aluminum alkyl on the polymerization and monomer-isomerization were observed; the polymer yield decreased in the following order: (CH3)3Al > (i–C4H9)3Al > (C2H5)3Al. Isomerization of VCHA was observed with the TiCl3–(i–C4H9)3Al and the TiCl3–(C2H5)3Al catalysts during the polymerization, while with the TiCl3–(CH3)3Al catalyst such isomerization was not observed. Monomer-isomerization copolymerization of VCHA and trans-2-butene took place to give copolymers consisting of VCHA and 1-butene units.  相似文献   

11.
Six alkyl alcohols were studied using thermospray mass Spectrometry. Whereas the dominant ion in the spectrum up to a repeller potential of 120 V was [M + NH4]+, above that potential [M + H]+ and fragment ions appeared. The fragments observed were largely due to hydrogen release from alkyl ions ([CnH2n+1]+ – H2 → [CnH2n-1]+) and loss of water or some other stable molecule from the same species. The results are compared with those from ionization of the same alcohols under electron impact and photoionization conditions and with results obtained for methanol under thermospray conditions.  相似文献   

12.
Polymerization of 1-methylthio-1-alkynes (MeSC?CR; R = Et, n-Bu, n-C6H13, and n-C8H17) was studied by use of transition metal catalysts. A 1 : 2 mixture of MoCl5 and Ph3SiH provided polymers having M?w over 1 × 105 in 30–50% yields from these monomers. The length of the alkyl group hardly affected the polymerization. The monomer, MeSC?C-n-C6H13, showed low reactivity in homopolymerization, but higher reactivity than that of MeC?C-n-C5H11 in copolymerization. Poly(1-methylthio-1-alkyne)s were colorless solids, and those with long alkyl pendants (R = n-C6H13, n-C8H17) were soluble in various organic solvents. The present polymers were thermally more stable than poly(2-alkyne)s, the corresponding hydrocarbon polymers.  相似文献   

13.

Abstract  

This work describes the regioselective synthesis of two new series of 1,1′-oxalylbis[3-(alkyl/aryl/heteroaryl)-4,5-dihydro-5-hydroxy-5-(trihalomethyl)-1H-pyrazoles], where the 3-substituents are H, Me, C6H5, 4-FC6H4, 4-ClC6H4, 4-BrC6H4, 4-NO2C6H4, 4,4′-BiPh, and 2-furyl, in a one-pot methodology with ethanol as solvent, from the reaction of 4-alkoxy-4-(alkyl/aryl/heteroaryl)-1,1,1-trihaloalk-3-en-2-ones with oxalyldihydrazide (51–89%). Complementarily, the dehydration reactions of five examples of the described oxalylbispyrazolines are also reported, which furnished the respective 1,1′-oxalylbis[3-(alkyl/aryl/heteroaryl)-5-(trihalomethyl)-1H-pyrazoles] in 53–78% yields without the two C(O)–N bond cleavages.  相似文献   

14.
A simple expression that allows accurate calculation of physicochemical properties of organic compounds like RX (R is an unbranched alkyl substituent C n H2 n + 1 and X is a functional group) was proposed. The potential of the proposed approach is demonstrated by the estimation of the boiling points and heat capacities at constant pressure of alkyl benzenes and carboxylic esters.  相似文献   

15.
New diphenyldiacetylenes of the type with A, B = H and/or F; m = 0, 1; n = 1–4; and X = C n H2n+1, F, CF3 or CN were synthesized and their mesomorphic properties determined by hot stage polarizing microscopy and DSC. When m = 0, all of these compounds showed only a nematic phase except when X = CF3 when both nematic and smectic A phases were seen. Both clearing and melting temperatures were higher than those reported for substitution with the corresponding alkyl chains but the much larger increase in clearing temperatures produced considerably wider nematic phases. Eutectic mixtures of a few of these olefins yielded nematic materials also having much wider temperature ranges and higher clearing temperatures than the eutectic mixtures of the alkyl compounds, while retaining their high birefringence and low viscosities. Such materials are of interest for beam-steering devices. Four of the diacetylenes with m = 1 (A, B = H) were also prepared (X = C6H13, F, n= 2, 3). When X was C6H13 (n=2), the nematic range was smaller in the 2- than in the 1-olefin but wider than in the alkyl series. When X=F, either no nematic phase or a monotropic one was observed, whereas the 1-olefins gave a much wider nematic phase. Both transition temperatures were lower than those for the corresponding 1-olefin and alkyl analogues. The compound with X=C6H13 and n=2 had a melting temperature below room temperature.  相似文献   

16.
Methyl radical reactions with matrix molecules in glasses C2H5OH, (CH2OH)2, n- and i-C3H7OH, n- and i-C4H9OH, n- and i-C5H11OH, C2D5OH, and i-C3D7OD, and the reactions of ?2H5, ?3H7, ?4H9, ?5H11 with methanol glasses have been studied. Alkyl radicals were produced by photolysis of diphenylamine–alkylhalide–alcohol mixtures using ultraviolet light. In all cases the alkyl radical decay follows the law c = c0 exp(-kt). The √t law should not be associated with alkyl radical diffusion in a matrix. A method of processing the kinetics of those reactions in which one paramagnetic species changes into another with the total concentration being constant and the electron spin resonance spectra of both species overlapping, is described.  相似文献   

17.
The cationic azo-surfactants possessing different spacers and tail alkyl chain lengths have been synthesized by azocoupling ofp-alkylaniline orop-ethoxyaniline with phenol, followed by alkylation and quaternalization with dibromoalkane and trimethylamine, respectively. These surfactants showed a good solubility in water. A reversibletrans-cis isomerization of the azosurfactants by photoirradiation was assessed by UV-Vis absorption spectra. Due to a difference in HLB between thetrans- andcis-surfactants, the observed critical micelle concentration (CMC) values and the electric conductivity of the surfactant solution at above the CMC were significantly affected by the photoinducedtrans-cis isomerization. The azo-surfactants bearing moderate alkyl chain lengths such as surfactants 6 (R2=C2H4, R3=C4H9) and 9 (R2=C4H8, R3=C2H5) were found to be effective to achieve large CMC changes (3.6 mmol/L for 6 and 5.9 mmol/L for 9) by UV-light irradiation. The replacement of the tail chain species also affected the photoresponsive function. The surfactant 12, possessingp-ethoxy group as the tail chain, was found to form a stable micelle aggregation as compared with the structurally related surfactant 10 having ethyl unit as its tail group, but it exhibited a large CMC change (5.3 mmol/L) by UV-light irradiation.  相似文献   

18.
The zwitterion, formed from the reaction of an alkyl isocyanide and a dialkyl acetylenedicarboxylate, reacts with phenacyl halides in H2O to produce γ‐iminolactone derivatives in high yields. H2O helps to avoid the use of highly toxic and environmentally unfavorable solvents for this conversion.  相似文献   

19.
A new and facile method for the general preparation of 3‐alkoxy‐2,3‐dihydro‐1H‐isoindol‐1‐ones has been developed. Thus, the reaction of 2‐(azidomethyl)benzoates with NaH affords, after workup with H2O, 3‐alkoxy‐2,3‐dihydro‐1H‐isoindol‐1‐ones 2 . 2‐Substituted 3‐alkoxy‐2,3‐dihydro‐1H‐isoindol‐1‐ones 4 can be obtained by adding alkyl halides prior to workup with H2O.  相似文献   

20.
(o-Phenylenediamino)borylstannanes were newly synthesized to achieve radical boryl substitutions of a variety of alkyl radical precursors. Dehalogenative, deaminative, decharcogenative, and decarboxylative borylations proceeded in the presence of a radical initiator to give the corresponding organic boron compounds. Radical clock experiments and computational studies have provided insights into the mechanism of the homolytic substitution (SH2) of the borylstannanes with alkyl radical intermediates. DFT calculation disclosed that the phenylenediamino structure lowered the LUMO level including the vacant p-orbital on the boron atom to enhance the reactivity to alkyl radicals in SH2. Moreover, C(sp3)-H borylation of THF was accomplished using the triplet state of xanthone.  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