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.
On Chalcogenolates. 151. Studies on Derivatives of N-Thioformyl Dithiocarbamic Acid. 1. Synthesis and Properties of N-Thioformyl Dithiocarbamates The N-thioformyl dithiocarbamates M[S2C? NH? CS? H], where M = K, Rb, Cs, Tl, NH4, [N(nC4H9)4], Na[S2C? NH? CS? H] · 0.5 H2O, and Ba[S2C? NH? CS? H]2 · 3 HO? CH2? CH2? OCH3 have been prepared by use of partial different procedures. The compounds were characterized with chemical and thermal methods as well as by means of electron absorption, infrared, nuclear magnetic resonance (1H and 13C), and mass spectra. Attempts to synthesize N-thioformyl dithiocarbamic acid were not successful.  相似文献   

2.
A new type of ethoxylated double‐tail trisiloxane surfactants containing a propanetrioxy spacer of the general formula ROCH2CH(OR)CH2O(CH2CH2O)xCH3 [R = Me3SiOSiMe(CH2)3OSiMe3, x = 8.4, 12.9, 22] has been synthesized. Their structures were characterized by 1H‐NMR, 13C‐NMR and 29Si‐NMR spectroscopy. The critical micelle concentration (CMC) values of these double‐tail trisiloxane surfactants were at the level of 10−5 mol l−1, and the surface tension values of their aqueous solutions at CMC were in the range of 21‐24.9 mN m−1. Only the double‐tail trisiloxane surfactant with average ethoxy units of 8.4 ( 1P ) possesseda good spreading ability (SA) value. Its SA values of aqueous solutions (5.0 × 10−3 mol l−1) on parafilm and Ficus microcarpa leaf surfaces were more than 15 (within 10 min) and 13 (within 3 min), respectively. The trisiloxane surfactant 1P was also found to have the strongest hydrolysis resistant ability among all of the double‐tail trisiloxane surfactants prepared. Its aqueous solutions were stable for 130 days in an acidic environment (pH 4.0) and 59 days in an alkaline environment (pH 10.0) with surface tension values less than 23 mN m−1. It is suggested that this surfactant can be used as a wetting agent or spreading agent in certain extreme pH environments. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

3.
IR data of eight substituted flavanones and their isomeric hydroxychalkones have been recorded in order to assign the various absorption bands and to study the effect of substituents on C=C out-of-plane deformation (400–700 cm?1), C?H out-of-plane deformation (700–1000 cm?1), C?H in-plane deformation (1000–1300 cm?1), C?O stretch (≈1200 cm?1), OCH3 (1200–1300 cm?1), O?H deformation (1300–1400 cm?1), CH3 deformation (1300–1500 cm?1), benzene ring vibration (1400–1600 cm?1) and C=O stretch (≈1650 cm?1). The δC?H (ring A) in 2′,4′-dihydroxy-3-nitrochalkone appears at 826 cm?1 (s), while in the isomeric flavanone it shows up as three bands, viz., 807 (w), 833 (m) and 881 cm?1 (w). This difference principally arises due to the presence of the electron withdrawing nitro substituent. The C=O stretching vibration in flavanones appears at a higher frequency than in the corresponding hydroxychalkones. This is perhaps due to the lack of conjugation in the former class of compounds. Chloro substituents (ring B) in different positions exert differing effects on νC?O. These differences can be rationalized in terms of a field-effect exerted by the chlorine atom.  相似文献   

4.
The spreading behaviour of defined trimethylsilane‐based surfactants of general formula (CH3)3Si(CH2)6(OCH2CH2) nOCH3, n = 2–6, on five different solid surfaces at 21 °C has been investigated. Compounds bearing short diethylene and triethylene glycol hydrophiles do not spread. For the longer‐chained tetraethylene to hexaethylene glycol derivatives, the ability to spread depends on the surface energy. Rapid spreading is restricted to the slightly polar surface of 40 mN m−1 surface energy. Lower or higher surface energies considerably reduce the spreading rates. The phase behaviour of the solutions substantially influences the spreading process. The dispersed systems of the tetraethylene glycol derivative spread constantly over long time intervals. The dispersions of the pentaethylene glycol analogue are very close to the temperature for a transition into the one‐phase state. A retardation of the spreading process occurs after a few seconds. Micellar solutions of the hexaethylene glycol derivative either spread very slowly or stop spreading after a few seconds. The largest spreading areas and highest initial spreading rates were found for the 0.1 wt% solutions. Copyright © 1999 John Wiley & Sons, Ltd.  相似文献   

5.
Nucleophilic substitution reactions of iodine pentaflurode with a series of homologous bifunctional alcoholates ?O(CH2)nO? (n=2,3,4,5,6,12), a geminal dialcoholate CC?3 CH(O?)2 and a trifunctional alcoholate CH3C(CH2O?)3 protected by (CH3)3Si - groups are reported. Systems with short CH2 - chains (n<4) first form short lived species IF4[O(CH2)nO]X (X = SiMe3, IF4) which rearrange to mononuclear chelates IF3[O(CH2)nO] of high stability. Dialcoholates with long CH2-chains (n>4) behave as bridging ligands forming stable multinuclear compounds IF4[O(CH2)nO]IF4 and {IF3[O(CH2)nO]}m (m≥2). 1,4-Butanediolate is on the border line of the two systems.Products with greater substitution IF[O(CH2)nO]2 (n=2,3) and IF2(OCH2)3CCH3 are also characterized.The dependence of 19F-NMR-shifts on the nature and arrangement of ligands is discussed.  相似文献   

6.
The spreading behaviour of defined trisiloxane surfactants of general formula [(CH3)3SiO]2 CH3Si(CH2)3(OCH2CH2) nOCH3 (n = 3–9) on five different solid surfaces has been investigated. Maximum spreading areas and rates are found on non‐polar or slightly polar surfaces of 30 to 40 mN m−1 surface energy. Extremely low or high surface energies substantially reduce the spreading rates. On non‐polar surfaces rapid spreading is observed for 1 wt % solutions of the relatively short‐chained penta‐ and hexa‐ethylene glycol derivatives. On slightly polar surfaces dilute 0.1 wt % solutions of longer‐chained derivatives spread faster. This spreading pattern shift coincides with a change of the phase behaviour. Solutions of Silwet L77 do not prefer one specific surface, since 1 wt % solutions abruptly stop spreading after a few seconds and the maximum spreading rates are found for 0.1 wt % solutions. Therefore, Silwet L77 essentially belongs among the long‐chained derivatives. Copyright © 2000 John Wiley & Sons, Ltd.  相似文献   

7.
The nucleophilic substitution of the reactive chlorine atoms of the n-butylboron-capped hexachlorine-containing clathrochelate precursor with the thiol-terminated closo-decaborate (PPh4)2[B10H9OCH2CH2OCH2CH2SH] in the presence of triethylamine afforded the closo-decaborate-containing clathrochelate dodecaanion [Fe{[(B10H9OCH2CH2OCH2CH2S)2Gm]3-(BBun)2}]12?, which was isolated in the form of its tetraphenylphosphonium and sodium salts. The complexes obtained were characterized using elemental analysis, UV-VIS, 1H, 11B, and 13C{1H} NMR, and IR spectroscopies.  相似文献   

8.
The organosilicone surfactant Silwet L‐77® (L‐77), used as an agrochemical adjuvant, is a mixture comprised predominantly of [(CH3)3SiO]2? (CH3)Si? (CH2)3? (OCH2CH2)n? OCH3 oligomers (n = 3–16, average n ≈ 7.5). The commercially available L‐77 mixture was purified by reversed‐phase high‐performance liquid chromatography (HPLC) to obtain individual trisiloxane surfactant components. Pure oligomers (n = 3, 6 and 9) were also synthesized. Synthesis was achieved by hydrosilylation of monomeric ethoxylate monomethyl ether starting reagents. Pure hexa‐ and nona‐ethylene glycols were produced by condensation of smaller oligomers. Atmospheric‐pressure ionization mass spectrometry (MS) methods were used to characterize fully the commercial L‐77 product and synthesized or isolated components. The application of Fourier‐transform ion cyclotron resonance MS and online HPLC–electrospray ionization MS techniques to the analysis of this surfactant are described here. The application of these analytical techniques also enabled elucidation of the synthetic by‐products present in the commercial formulation. In addition, physico‐chemical properties specific to agrochemical uses, such as droplet spread areas on plant foliage and surface tension for the different oligomer solutions, are also reported. Copyright © 2004 John Wiley & Sons, Ltd.  相似文献   

9.
By combining results from a variety of mass spectrometric techniques (metastable ion, collisional activation, collision-induced dissociative ionization, neutralization-reionization spectrometry, 2H, 13C and 18O isotopic labelling and appearance energy measurements) and high-level ab initio molecular orbital calculations, the potential energy surface of the [CH5NO]+ ˙ system has been explored. The calculations show that at least nine stable isomers exist. These include the conventional species [CH3ONH2]+ ˙ and [HO? CH2? NH2]+ ˙, the distonic ions [O? CH2? NH3]+ ˙, [O? NH2? CH3]+ ˙, [CH2? O(H)? NH2]+ ˙, [HO? NH2? CH2]+ ˙, and the ion-dipole complex CH2?NH2+ …? OH˙. Surprisingly the distonic ion [CH2? O? NH3]+ ˙ was found not to be a stable species but to dissociate spontaneously to CH2?O + NH3+ ˙. The most stable isomer is the hydrogen-bridged radical cation [H? C?O …? H …? NH3]+ ˙ which is best viewed as an immonium cation interacting with the formyl dipole. The related species [CH2?O …? H …? NH2]+ ˙, in which an ammonium radical cation interacts with the formaldehyde dipole is also a very stable ion. It is generated by loss of CO from ionized methyl carbamate, H2N? C(?O)? OCH3 and the proposed mechanism involves a 1,4-H shift followed by intramolecular ‘dictation’ and CO extrusion. The [CH2?O …? H …? NH2]+ ˙ product ions fragment exothermically, but via a barrier, to NH4+ ˙ HCO…? and to H3N? C(H)?O+ ˙ H˙. Metastable ions [CH3ONH2]+…? dissociate, via a large barrier, to CH2?O + NH3+ + and to [CH2NH2]+ + OH˙ but not to CH2?O+ ˙ + NH3. The former reaction proceeds via a 1,3-H shift after which dissociation takes place immediately. Loss of OH˙ proceeds formally via a 1,2-CH3 shift to produce excited [O? NH2? CH3]+ ˙, which rearranges to excited [HO? NH2? CH2]+ ˙ via a 1,3-H shift after which dissociation follows.  相似文献   

10.
Detailed normal-coordinate analysis has been carried out on a large number of conformers of model molecules of poly(oxyethylene); the model molecules treated are CH3(OCH2CH2)nOCH3 with n = 2,3, and 6. The systematic treatment provides well-defined correlations between conformation and vibrational spectra of poly(oxyethylene). The vibrations in the region 1050–800 cm?1, which are associated with the C? O and C? C stretching and CH2 rocking modes, are highly dependent on the conformation of the polymer chain. On the basis of these correlations, the infrared and Raman spectra of the molten state and of the aqueous solution are interpreted in terms of the conformational states. The analysis indicates that the following conformational fragments are present in these phases: GT-TG (T: trans; G: gauche), TT-TG, GT-GG, TT-TT, and TT-GG for the CH2? CH2? O? CH2? CH2 group, and TGT and TGG for the O? CH2? CH2? O group. Conformational fragments GT-TG′ and GT-GG′ are also possible in the analogy to GT-TG and GT-GG.  相似文献   

11.
Abstract

Electrophilic trisubstituted ethylenes, ring‐substituted ethyl 2‐cyano‐3‐phenyl‐2‐propenoates, RC6H4CH?C(CN)CO2C2H5 (where R is 2‐CH3, 3‐CH3, 4‐CH3, 2‐OCH3, 3‐OCH3, and 4‐OCH3) were prepared and copolymerized with styrene (ST). The monomers were synthesized by the piperidine catalyzed Knoevenagel condensation of ring‐substituted benzaldehydes and ethyl cyanoacetate, and characterized by CHN analysis, IR, 1H and 13C NMR. All the ethylenes were copolymerized with ST (M1) in solution with radical initiation (AIBN) at 70°C. The compositions of the copolymers were calculated from nitrogen analysis and the structures were analyzed by IR, 1H and 13C NMR. The order of relative reactivity (1/r 1) for the monomers is 3‐OCH3 (0.88)?>?4‐CH3 (0.71)?>?2‐OCH3 (0.68)?>?3‐CH3 (0.55)?>?2‐CH3 (0.47)?>?4‐OCH3 (0.40). Higher T g of the copolymers in comparison with that of polystyrene indicates a decrease in chain mobility of the copolymer due to the high dipolar character of the TSE structural unit. Gravimetric analysis indicated that the copolymers decompose in the 257–287°C range.  相似文献   

12.
Triorganotin(IV) pyrazolinates of the type R3Sn(C15H12N2O?·?X) [where C15H12N2O?·?X?= 3(2′-hydroxyphenyl)-5(4-X-phenyl)pyrazoline {where X?=?H (a); CH3 (b); OCH3 (c); Cl (d) and R?=?Me, Pr n and Ph}] have been synthesized by the reaction of R3SnCl with the sodium salt of pyrazolines in a 1?:?1 molar ratio, in anhydrous benzene. These newly synthesized derivatives have been characterized by elemental analysis (C, H, N, Cl and Sn), molecular weight measurement as well as spectral studies [IR and multinuclear NMR (1H, 13C and 119Sn)]. The bidentate behaviour of the ligands was confirmed by IR, 1H and 13C NMR spectral data. Trigonal bipyramidal structure around tin(IV) atom for R3Sn(C15H12N2O?·?X) has been suggested. The free pyrazoline and a few triorganotin(IV) pyrazolinates have also been screened for their antibacterial and antifungal activities. Some triorganotin(IV) pyrazolinates exhibit higher antibacterial and antifungal effects than free pyrazoline and some of the antibiotics.  相似文献   

13.
IntroductionInrecentyears ,bis(quaternaryammonium)surfac tantsorgeminisurfactants ,inwhichtwocationicsurfac tantmoietiesareconnectedwiththeammoniumheadgroupbyaploymethylenechain ,namely ,aspacerhavebecomeofinterestduetotheirexceptionalsurfaceactivityandrem…  相似文献   

14.
Chlorodiorganotin(IV)pyrazolinates of the type R2SnCl(C15H12N2O?·?X) [where C15H12N2O?·?X?=?3(2′-Hydroxyphenyl)-5(4-X-phenyl)pyrazoline {where X?=?H (a); CH3 (b); OCH3 (c); Cl (d) and R?=?Me, Pr n and Ph}] have been synthesized by reaction of R2SnCl2 with the sodium salt of pyrazolines in 1?:?1 molar ratio, in anhydrous benzene. These newly synthesized derivatives have been characterized by elemental analysis (C, H, N, Cl and Sn), molecular weight measurement and spectral studies [IR and multinuclear NMR (1H, 13C and 119Sn)]. The bidentate behavior of the ligands was confirmed by IR, 1H and 13C NMR spectral data. Trigonal bipyramidal structure around tin(IV) for R2SnCl(C15H12N2O?·?X) is suggested. The free pyrazoline and some chlorodiorganotin(IV) pyrazolinates have been screened for their antibacterial and antifungal activities. Some chlorodiorganotin(IV) pyrazolinates exhibit higher antibacterial and antifungal effect than free pyrazoline and some antibiotics.  相似文献   

15.
Positive and negative cluster ions in methanol have been examined using a direct fast atom bombardment (FAB) probe technique. Positive ion (CH3OH)IIH + clusters with n = 1-28 have been observed and their clusters are the dominant ions in the low-mass region. Cluster-ion reaction products (CH3OH)II(H2O)H+ and (CH3OH)II(CH3OCH3)H+ are observed for a wide range of n and the abundances of these ions decrease with increasing n. The negative ion (CH3OH)II(CH3O)? clusters are also readily observed with n = 0-24 and these form the most-abundant negative ion series at low n. The (CH3OH)II(CH2O)?, (CH3OH)II(HIIO)(CH2O)? and (CH3OH)II(H2OXCH3O)? cluster ions are formed and the abundances of these ions approach those of the (CH3OH)II(CH3O)? ion series at high n. Cluster-ion structures and energetics have been examined using semi-empirical molecular orbital methods.  相似文献   

16.
Complexes of iron(III) with ethylene glycol and 3(2′-hydroxyphenyl)-5-(4′-substituted phenyl) pyrazolines, [Fe(OCH2CH2O)(C15H12N2OX)] m ? nH2O and [Fe(C15H12N2OX)2(OCH2CH2OH)] (where OCH2CH2O and OCH2CH2OH = ethylene glycol moiety; C15H12N2OX = 3(2′-hydroxyphenyl)-5-(4-X-phenyl)pyrazoline; X = H, CH3, OCH3, or Cl; m = 2–3 and n = 2–3) have been synthesized and characterized by elemental analysis (C, H, N, Cl, and Fe), molecular weight measurement, magnetic moment data, thermogravimetric analysis, molar conductance, spectral (UV-Vis, IR, and FAB mass), scanning electron microscopy, and X-ray diffraction studies. Bonding of ethylene glycol and pyrazolines in these complexes and the particle size of iron(III) complexes are discussed. Antibacterial and antifungal potential of free pyrazoline and some iron(III) complexes are also discussed.  相似文献   

17.
Template combination of copper acetate (Cu(AcO)2?H2O) with sodium dicyanamide (NaN(C≡N)2, 2 equiv) or cyanoguanidine (N≡CNHC(=NH)NH2, 2 equiv) and an alcohol ROH (used also as solvent) leads to the neutral copper(II)–(2,4‐alkoxy‐1,3,5‐triazapentadienato) complexes [Cu{NH?C(OR)NC(OR)?NH}2] (R=Me ( 1 ), Et ( 2 ), nPr ( 3 ), iPr ( 4 ), CH2CH2OCH3 ( 5 )) or cationic copper(II)–(2‐alkoxy‐4‐amino‐1,3,5‐triazapentadiene) complexes [Cu{NH?C(OR)NHC(NH2)?NH}2](AcO)2 (R=Me ( 6 ), Et ( 7 ), nPr ( 8 ), nBu ( 9 ), CH2CH2OCH3 ( 10 )), respectively. Several intermediates of this reaction were isolated and a pathway was proposed. The deprotonation of 6 – 10 with NaOH allows their transformation to the corresponding neutral triazapentadienates [Cu{NH?C(OR)NC(NH2)?NH}2] 11 – 15 . Reaction of 11 , 12 or 15 with acetyl acetone (MeC(?O)CH2C(?O)Me) leads to liberation of the corresponding pyrimidines NC(Me)CHC(Me)NC NHC(?NH)OR, whereas the same treatment of the cationic complexes 6 , 7 or 10 allows the corresponding metal‐free triazapentadiene salts {NH2C(OR)?NC(NH2)?NH2}(OAc) to be isolated. The alkoxy‐1,3,5‐triazapentadiene/ato copper(II) complexes have been applied as efficient catalysts for the TEMPO radical‐mediated mild aerobic oxidation of alcohols to the corresponding aldehydes (molar yields of aldehydes of up to 100 % with >99 % selectivity) and for the solvent‐free microwave‐assisted synthesis of ketones from secondary alcohols with tert‐butylhydroperoxide as oxidant (yields of up to 97 %, turnover numbers of up to 485 and turnover frequencies of up to 1170 h?1).  相似文献   

18.
Hydrosilylation of fluorinated olefins with polyhydromethylsiloxane (PHMS) in the presence of a platinum catalyst was investigated to synthesize fluorosilicone having highly fluorinated alkyl side chains (Rf; CnF2n+1? ). The hydrosilylation of 3,3,4,4,5,5,6,6,7,7,8,8,9,9,10,10,10‐heptadecafluoro‐1‐decene (C8F17CH?CH2) ( 1 ) with poly(dimethylsiloxane‐co‐hydromethylsiloxane) {(CH3)3SiO[? (H)CH3SiO? ]8[? (CH3)2 SiO? ]18Si(CH3)3} ( 4 ) converted the hydrogen bonded to silicons into the 3,3,4,4,5,5,6,6,7,7,8,8,9,9,10,10,10‐heptadecafluorodecyl group or fluorine bonded to silicons in the ratio of about 52:48, and the formation of the byproduct C7F15CF?CHCH3 ( 8 ) was observed. The hydrosilylation of 7,7,8,8,9,9,10,10,11,11,12,12,13,13,14,14,14‐heptadecafluoro‐4‐oxa‐1‐tetradecene (C8F17CH2CH2OCH2CH?CH2) ( 2 ) with 4 converted the hydrogen bonded to silicons into the 7,7,8,8,9,9,10,10,11,11,12,12,13,13,14,14,14‐heptadecafluoro‐4‐oxa‐tetradocyl group bonded to silicons, but an excess amount of 2 was required to complete the reaction because the isomerization of 2 occurred in part to form C8F17CH2CH2OCH?CHCH3 ( 9 ). The hydrosilylation of 4,4,5,5,6,6,7,7,8,8,9,9, 10,10,11,11,11‐heptadecafluoro‐1‐undecene (C8F17CH2CH?CH2) ( 3 ) with 4 converted the hydrogen bonded to silicons into the 4,4,5,5,6,6,7,7,8,8,9,9,10,10,11,11‐heptadecafluoroundecyl group bonded to silicons. This type of fluorinated olefin was successfully applied to the hydrosilylation with other PHMS's that involved a homopolymer of PHMS and a cyclic PHMS. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 3120–3128, 2002  相似文献   

19.
Abstract

A series of α-substituted selenenyl acetophenone derivatives of the types, [PhC(OCH2CH2O)CH2Se]2, [PhC(OCH2CH2O)CH2SeR], (PhCOCH2Se)2, and [PhCOCH2SeR] have been prepared. These compounds have been characterized by elemental analyses, IR and NMR (1H, 13C, 77Se) spectroscopy. The compounds, [PhC(OCH2CH2)CH2Se]2 and (PhCOCH2Se)2 have been structurally characterized by single crystal X-ray diffraction analyses. The former shows intra-molecular Se‐?‐?‐O interaction, while the latter exhibits inter-molecular nonbonding Se‐?‐?‐O interaction.

[Supplementary materials are available for this article. Go to the publisher's online edition of Phosphorus, Sulfur, and Silicon and the Related Elements for the following free supplemental files: Additional text and figures.]  相似文献   

20.
By combining results from a variety of mass spectrometric techniques (metastatle ion, collisional activation, collision-induced dissociative ionization, neutralization–reionization spectrometry and appearance energy measurements) and the classical method of isotopic labelling, a unified mechanism is proposed for the complex unimolecular chemistry of ionized 1,2-propanediol. The key intermediates involved are the stable hydrogen-bridged radical cations [CH2?C(H)? H…?O…?O(H)CH3]+˙, which were generated independently from [4-methoxy, 1-butanol]+˙ (loss of C2H4) and [1-methoxyglycerol]+˙ (loss of CH2O), [CH3? C?O…?H…?O(H)CH3]+˙ and the related ion-dipole complex [CH2?C(OH)CH3/H2O]+˙. The latter species serves as the precursor for the loss of CH3˙ and in this reaction the same non-ergodic behaviour is observed as in the loss of CH3˙ from the ionized enol of acetone.  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