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 46 毫秒
1.
A study of three isomeric compounds containing a phenolic moiety attached to the nitrogen‐rich triazolo–triazole bicycle is presented. In the three isomers, the phenolic OH group is in the ortho, meta and para positions. The crystal structure analysis of the meta isomer (C10H9N5O) shows that the 2H‐tautomer is present in the crystal and that the molecule adopts a substantially planar geometry. However, the conformation found in the crystal is different compared to the monoprotonated cation of the same compound previously investigated in several salts. The packing of the meta isomer is driven by the formation of strong hydrogen bonds and shows the formation of infinite planar ribbons, parallel to a, formed around 21 crystallographic axes. The three isomers were tested against some cancer cell lines and also against normal cell lines. The ortho isomer shows a weak antiproliferative activity, the meta isomer shows significant antiproliferative activity against some cancer lines and no activity against healthy cell lines, and the para isomer is active against all the tested cell lines.  相似文献   

2.
The recent introduction of multiangle light-scattering detectors for size-exclusion chromatography has made possible the measurement of the root mean square radius of gyration (Rg) and molecular weight (M) of polymer fractions eluting from a size-exclusion chromatography column. The characterization of the dimensions of a polymer may be accomplished with only a few milligrams. The dimensions of a polyimide precursor prepared by the condensation of the meta-diethyl ester of pyromellitic dianhydride with para-phenylene diamine have been measured with this technique. The dependence of Rg on M across the distribution is compared with that predicted for a freely rotating chain, and with other similar polymers measured with hydrodynamic techniques.  相似文献   

3.
The kinetics of the dehydrogenation of the individual ortho, meta, and para isomers of perhydroterphenyl and their mixtures over a (3 wt % Pt)/C catalyst has been investigated in a flow reactor at 280–340°C. The rate of the isomerization of the stereoisomers of the initial substrate (perhydroterphenyl) and terphenyl dehydrogenation products has an effect on the hydrogen release kinetics. The highest reactivity in isomerization is shown by the ortho isomer. The largest amount of hydrogen (7.0 wt %) is released in the dehy-drogenation of perhydro-meta-terphenyl and perhydro-para-terphenyl, whose conversion at 320°C is 96%.  相似文献   

4.
Copolymerizations of three phthalaldehyde isomers (M2) with styrene (M1) were carried out in methylene chloride or in toluene with BF3OEt2 catalyst. The monomer reactivity ratios were r1 = 0.77, r2 = 0 for the meta isomer and r1 = 0.60, r2 = 0 for the para isomer. The second aldehyde group of both isomers did not participate in polymerization and acted simply as the electron-withdrawing group, thus reducing the cationic reactivity of these monomers. Copolymerization behaviors of the ortho isomer (o-PhA) were quite different between 0°C and ?78°C. At ?78°C, o-PhA preferentially polymerized to yield “living” cyclopolymers, until an equilibrium concentration of o-PhA monomer was reached. Then, styrene propagated from the living terminal rather slowly. The block structure of the copolymer was confirmed by the chemical and spectroscopic means. In the copolymerization at 0°C, the o-PhA unit in copolymer consisted both of cyclized and uncyclized units. This copolymer seemed to contain short o-PhA sequences. The variation of the o-PhA-St copolymer structure with the polymerization temperature was explained on the basis of whether the polymerization was carried out above or below the ceiling temperature (?43°C) of the homopolymerization of o-PhA.  相似文献   

5.
Radical polymerization of isopropyl tert-butyl fumarate (iPtBF) and monomer-isomerization radical polymerization of isopropyl tert-butyl maleate (iPtBM) were investigated with 2,2′-azobisisobutyronitrile as initiator in the presence and absence of morpholine (Mor) as isomerization catalyst. It was found that iPtBF gave high molecular weight polymers in high yield as previously observed for diisopropyl fumarate (DiPF) and di-tert-butyl fumarate (DtBF). It was confirmed that iPtBF produced by in situ monomer isomerization of iPtBM homopolymerized to give a polymer. Radical copolymerization of iPtBM with styrene in the presence and absence of Mor was also performed and monomer reactivity ratios obtained were compared. From the kinetic study of the isomerization of iPtBM, it was revealed that the isomerization rate showed first-order dependence on the concentration of iPtBM and Mor, and that the apparent activation energy was 29.4 kJ/mol. On pyrolysis of the poly(iPtBF) at 180°C, isobutene and isopropanol were eliminated rapidly to yield polymer containing carboxyl groups and anhydrides. The pyrolytic behavior was different from that of a copolymer of DiPF with DtBF.  相似文献   

6.
An alkyl‐radical loss and an alkene loss are two competitive fragmentation pathways that deprotonated aminobenzoate esters undergo upon activation under mass spectrometric conditions. For the meta and para isomers, the alkyl‐radical loss by a homolytic cleavage of the alkyl‐oxygen bond of the ester moiety is the predominant fragmentation pathway, while the contribution from the alkene elimination by a heterolytic pathway is less significant. In contrast, owing to a pronounced charge‐mediated ortho effect, the alkene loss becomes the predominant pathway for the ortho isomers of ethyl and higher esters. Results from isotope‐labeled compounds confirmed that the alkene loss proceeds by a specific γ‐hydrogen transfer mechanism that resembles the McLafferty rearrangement for radical cations. Even for the para compounds, if the alkoxide moiety bears structural motifs required for the elimination of a more stable alkene molecule, the heterolytic pathway becomes the predominant pathway. For example, in the spectrum of deprotonated 2‐phenylethyl 4‐aminobenzoate, m/z 136 peak is the base peak because the alkene eliminated is styrene. Owing to the fact that all deprotonated aminobenzoate esters, irrespective of the size of the alkoxy group, upon activation fragment to form an m/z 135 ion, aminobenzoate esters in mixtures can be quantified by precursor ion discovery mass spectrometric experiments. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

7.
Three cyanate esters containing phosphorus are synthesized in good overall yields starting from bromoanisoles. Di‐ and tricyanates with meta configuration are most stable while para is less so. The para dicyanate ester isomer is particularly affected by water from the atmosphere. The meta dicyanate ester 2 has good thermal properties with glass transition at 268 °C and char yield of 65% in air at 600 °C. All three phosphorus‐containing cyanate esters are low flammability in an open flame. They make highly combustible cyanate esters resins less flammable simply by blending. Mixing 10 wt% dicyanate ester 2 into bisphenol A or E dicyanate esters makes them rate V‐0. Published 2018. J. Polym. Sci., Part A: Polym. Chem. 2018 , 56, 1100–1110  相似文献   

8.
Flash vacuum pyrolysis (FVP) of tert-butylthiosulfinic acid S-tert-butylester, t-BuS(O)St-Bu, at a temperature of 500 °C and a pressure of 0.07 hPa leads to the formation of tert-butylthiosulfoxylic acid, t-BuSSOH ( 1 ), and 2-methylpropene as byproduct. 1 has been identified in the gas phase by its IR absorption bands at ν(OH) = 3598 cm–1, δ(SOH) = 1149 cm–1 and ν(SO) = 718 cm–1. At higher temperatures (700 °C) the elimination of a second mole of 2-methylpropene and the shift of ν(SO) to higher wavenumbers (750 cm–1) indicate the formation of 1-oxatrisulfane, HSSOH. Different sulfenic acids RSOH (R = Me, i-Pr, t-Bu) were synthesized by FVP in order to study the influence of the substituent R on the vibrational wavenumbers ν(OH), ν(SO) and δ(SOH) observed in the gas phase. The strongest effect results for δ(SOH) leading to a decrease by 6 wavenumbers if the methyl group is substituted by a tert-butyl group. The assignment of the experimental wavenumbers has been supported by theoretical values obtained from ab initio calculations at the MP2(fc)/6-311G** level. Furthermore, the theoretical studies show that of all compounds RS2OR′ (R = R′ = H, Me; R = Me (H), R′ = H (Me)) the unbranched chain isomers RSSOR′ are energetically favored over the branched chain isomers. Relaxed potential energy surface scans at the MP2(fc)/6-311G** level have been carried out to study the rotational isomers of the branched molecules RS(Y)XR′ (R = R′ = H, Me; R = Me (H), R′ = H (Me); X = O (S), Y = S (O)). Of the three conformations (+)syn-clinal, (–)syn-clinal, and anti-periplanar resulting from molecular model considerations only the two latter ones correspond to local minima on the calculated potential curve. The (–)syn-clinal conformation is slightly favored for all other constitutional isomers except for HS(O)SH and MeS(O)SH, which prefer the anti-periplanar conformation.  相似文献   

9.
Treatment of 2-(tert-butyl)-1,2,3,4-benzotetrazinium tetrafluoroborates with sodium thiocyanate afforded 2-(tert-butylazo)phenyl isothiocyanates 3, which exist in equilibrium with 2-(tert-butyl)-1,2,4-benzotriazine-3(2H)-thiones 3′. The equilibrium depends on the substituents R in the benzene ring: the percentage of the open isomer 3 is about 20% for R = H or Me; for R = Cl or Br, the equilibrium is completely shifted to cyclic isomer 3′. The equilibrium is slow on the time scale of the 1H and 13C NMR experiments. For compounds 3a/3′a (R = H), the spectra at 24 °C show two sets of signals, while those at 0 °C contain only signals for isomer 3′a. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 7, pp. 1192–1195, July, 2006.  相似文献   

10.
4,4′‐Methylenebis(2‐tert‐butylaniline) was synthesized and reacted with pyromellitic dianhydride to produce a polyimide that showed excellent solubility in conventional organic solvents. Solutions of this polyimide could be cast into transparent, flexible and tough films. The number‐average molecular weight, as determined by means of gel permeation chromatography, was 8.9×104 g/mol and the polydispersity index was 1.97. The glass transition temperature was found to be 217°C. The polyimide did not show appreciable decomposition up to 500°C under a nitrogen atmosphere.  相似文献   

11.
Element-Element Bonds. I. Syntheses and Structure of Tetra(tert-butyl)tetrarsetane and of Tetra(tert-butyl)tetrastibetane Dilithium (tert-butyl)arsenide reacts with (tert-butyl)dichloroarsine to give tetra-(tert-butyl)tetrarsetane 1 ; homologous tetra(tert-butyl)tetrastibetane 2 is formed by reduction of (tert-butyl)dichlorostibane with magnesium. The isotypic compounds 1/2 crystallize in the monoclinic space group P21/c with Z = 4. The dimensions of the unit cells determined at ?45 ± 5°C are: a = 957.4(8)/1 000.2(3); b = 1 399.1(14)/1 423.9(4); c = 1 697.4(9)/1 749.8(7) pm; β = 96.02(6)/96.77(3)°. As shown by low temperature X-ray structure determinations (3 531/3 232 symmetry independent reflections; Rg = 4.0/4.6%) the four membered rings E4 (E = As or Sb) are folded; in all-trans configuration the bulky organic substituents occupy pseudo-equatorial positions. Characteristic averaged bond distances and angles are: E? E 244/282; E? C 202/221 pm; ? E? E? E 86/85° ? E? E? C 101/99°. The dihedral angels of the bisphenoides built up by the atoms of the rings are found to be 139/133°.  相似文献   

12.
Enantioselective host-guest complexes of α-amino-alcohol salts with chiral tartaric-acid esters can not be crystallised up to now. To study structural aspects of their enantioselectivity, crystal structures of the components were determined. The structures of eleven diesters with myrtanol, borneol, menthol, neomenthol, and cis-4-(tert-butyl)cyclohexanol in different configurations showed a remarkable rigidity of the tartaric-acid conformation, partly because of intramolecular H-bonding between OH and C=O groups. The conformation of the tartaric-acid part in these diesters is the same as the one observed in optically active tartaric acid (torsion angle OC? C? OH ca. 0°), The binding site for guest molecules is a parallelogram formed by two hydroxy and two carbonyl O-atoms, all lying on the same side of a mean molecular plane. There is one exception: the dimenthyl ester, which is the most enantioselective with a norephedrine guest, has one of the ester groups turned (torsion angle O? C? C? OH ca. 180°), forming a triangle of O-atoms and moving the bulky menthyl group to the vicinity of the binding site.  相似文献   

13.
Two diamines, 2,5-bis (4-aminophenyl)-2,5-diazahexane and 1,4-bis (4-aminophenyl)-1,4-diazacyclohexane were chosen as components for polyimidizations because they have melting points that differ by nearly 200°C (66–67 and 229–230°C, respectively) and are relatives of p-nitro-N,N-dimethylaniline. The melting points of the model compounds (phthalic anhydride) do not differ by as much as those of the free amines [303–304 and 386°C (DSC), respectively]. Six polyimides were prepared by a two-step polycondensation of the diamines with pyromellitic dianhydride, benzophenonetetracarboxylic dianhydride, and 5,5'-[2,2,2-trifluoro-1-(trifluoromethyl) ethylidene] bis-1,3-isobenzofurandione. DSC thermograms failed to indicate any distinct transitions up to 450°C, however, the polyimide prepared from 2,5-bis (4-aminophenyl)-2,5-diazahexane and pyromellitic dianhydride shows a slight break in its DSC curve at 233°C.  相似文献   

14.
Reaction of vanadyl acetylacetonate with tert-butyl hydroperoxide (benzene, 20°C) at any molar ratio leads to the elimination of ligand and its oxidation mainly to CO2 and acetic acid. At the (acac)2VO: t-BuOOH ratio above 1:10 liberation of oxygen partially in the singlet state takes place.  相似文献   

15.
Methods for the calculation of the torsional angle of the C? C linkage
  • 1 Voir Réf. 1.
  • are applied to some 4-germa-1,3-dioxanes. It is thereby shown that 4,4-diethyl-2trichloromethyl-4-germa-1,3-dioxane in CCl 4 and C 6 D 6 adopts the chair conformation, with the equatorial C? Cl 3 group distorted by the presence of the germanium: the torsional angle of the Ge-CH 2 -CH 2 -O-fragment is 45°. The most stable 6-alkyl derivatives ( cis isomers) have the same conformation; the less stable trans - tert -butyl isomer prefers a skew-boat form compatible with a torsional angle of about 60°.  相似文献   

    16.
    Synthesis and Molecular Structure of (N,N′-Dimethyl-piperazine)lithium-(·-hydrido)(tert-butyl)bis[bis(trimethylsilyl)methyl]alanate with an Intramolecular Interaction between Lithium and C? H-σ-Bonds Syntheses and properties of the starting compounds bis[bromo-di(tert-butyl)alane] 3 , bis[dibromo-tert-butyl-alane] 4 , and (tert-butyl)bis[bis(trimethylsilyl)methyl]alane 5 are described. In the presence of 5 and the chelating amine N,N′-dimethylpiperazine lithium tert-butyl gives via μ-elimination isobutene and LiH, which is taken up by the starting alane 5 to give the title compound 6 . No attack of the strong base (lithium alkyl/amine) to the bis(trimethylsilyl) methyl substituent is observed as recently occured for the sterically more crowded tris[bis(trimethylsilyl)methyl]alane. Crystal structure of 6 shows a angled Li? H? Al bridge and a short intramolecular contact between Li and C? H-σ-bonds of a trimethylsilyl group.  相似文献   

    17.
    Poly(amic acid) was synthesized with a low‐temperature solution polymerization of 3,3′‐dihydroxybenzidine and pyromellitic dianhydride in N,N‐dimethylacetamide. The cast films were thermally treated at various temperatures. The polyimide containing the hydroxyl group was rearranged by decarboxylation, resulting in a fully aromatic polybenzoxazole at temperatures higher than 430 °C. These stepwise cyclizations were monitored with elemental analysis, Fourier transform infrared, and nuclear magnetic resonance. Microanalysis results confirmed the chemical compositions of poly(amic acid), polyimide, and polybenzoxazole, respectively. A cyclodehydration from poly(amic acid) to polyimide occurred between 150 and 250 °C in differential scanning calorimetry, and a cyclodecarboxylation to polybenzoxazole appeared at 400–500 °C. All the samples were stable up to 625 °C in nitrogen and displayed excellent thermal stability. Polybenzoxazole showed better thermal stability than polyimide, but polyimide exhibited better mechanical properties than polybenzoxazole. However, polyimide showed a crystalline pattern under a wide‐angle X‐ray, whereas polybenzoxazole was amorphous. The precursor poly(amic acid) was readily soluble in a variety of solvents, whereas the polyimide and polybenzoxazole were not soluble at all. © 2000 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 38: 2537–2545, 2000  相似文献   

    18.
    A series of hexachloronadimides containing phosphine oxide in the backbone were synthesized by the reaction of bis(3-amino phenyl) methyl phosphine oxide (BAP) with pyromellitic dianhydride (PMDA)/3,3’,4,4’-benzophenone tetracarboxylic acid dianhydride (BTDA)/2,2-bis(3,4-dicarboxy phenyl) hexafluoropropane dianhydride (6F) and hexachloronadic anhydride in glacial acetic acid/acetone. Structural characterisation of the resins was carried out by infrared, nuclear magnetic resonance spectroscopy and elemental analysis. Thermal characterisation of uncured resin was done by differential scanning calorimetry and thermogravimetric analysis. The decomposition temperature of uncured resins were above 310±10°C with T max 330±10°C in nitrogen atmosphere. Char yield at 800°C ranged from 37–42%. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

    19.
    Herein we report, a rhodium-catalyzed Fujiwara-Moritani-type reaction of unactivated terminal alkenes and benzoic acid derivatives bearing electron donating residues under mild conditions. The acid functionality acts as a traceless directing group delivering products alkenylated in meta-position to the electron donating substituent in contrast to the usually obtained ortho- and para-substitution in Friedel-Crafts-type reactions. Remarkably, the new C−C bond is formed to the C2 of the terminal olefin, in contrast to similar reported transformations. Initially formed mixtures of exo- and endo-double bond isomers can be efficiently isomerized to the more stable endo-products.  相似文献   

    20.
    Di-tert-butyl maleate (DtBM) did not polymerize with 2,2′-azobis(isobutyronitrile) as a radical initiator, but DtBM easily homopolymerized via a monomer-isomerization radical polymerization mechanism to give a high molecular weight polymer when morpholine was added into the polymerization system as an isomerization catalyst. The feature of the monomer-isomerization polymerization of DtBM was investigated in detail. The polymer obtained was confirmed to consist of a poly(tert-butoxycarbonylmethylene) structure similar to that from di-tert-butyl fumarate. Subsequent pyrolysis of the resulting polymer at 180°C is a useful route to synthesis of a high molecular weight poly(fumaric acid). © 1993 John Wiley & Sons, Inc.  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