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 420 毫秒
1.
The ionic equilibria for poly-4-vinyl pyridine (P4VP) and poly-2-vinyl pyridine (P2VP) were studied by physico-chemical techniques such as potentiometry, viscosity and NMR-1H. The mixture of ethanol (45 per cent w.p.) and water was used as solvent to obtain the total range of ionization (0–1). It was found that the dissociation constants of pyridine residue of polymers in the absence of electrostatic interaction (pK0 = 3·3–3·9) are lower than for the monomer analogues 4-ethylpyridine and 2-ethylpyridine (5·02) and depend on ionic strength (NaCl).A sharp decrease of pKapp at the beginning of titration and increase of specific viscosity for P4VP and P2VP are both explained by electrostatic interactions between positive charges forming during titration of the macromolecules. Most probably, these interactions act through the organic part of the macromolecule. On the other hand, it is shown by NMR-1H that sharp changes in pKapp and specific viscosity at the beginning of the titration are not associated with changes in the average conformation of the monomer unit in the polymer. This conformation can be destroyed only when the energy of electrostatic interactions is large enough and this occurs when the mean distance between positive charges is relatively short.  相似文献   

2.
Water-soluble derivatives of poly-4-vinylpyridine (PVP) were prepared by means of quaternization of PVP with ethyl bromide, butyl bromide, hexyl-bromide and bromacetic acid. In addition, hydrolysis of p-nitrophenylacetate (NPA) and 3-nitro-4-acetoxybenzoic acid (NABA) catalyzed by these polymers and 4-ethylpyridine as analogue were studied spectrophotometrically both in water and alcohol-water mixtures at 25° and pH 7·8.The second-order rate of NPA hydrolysis (Kt) was independent of the length of alkyl groups and the salt concentration. While an increase of the degree of quaternization (β) for PVP derivatives resulted in a reduction of the value of Ki in water, in 37–50 vol per cent alcohol-water there was the opposite effect. At large values of β, Ki did not depend on the content of ethanol.The bell-shaped dependence of Ki on β was shown in the case of amionic NABA. The value of Ki proved to be very sensitive to both the nature of salt and the salt concentration in aqueous solution.The use of viscosity, turbidimetry and potentiometry for studying these polymers allowed the role of the quaternized pyridine residue to be elucidated. The peculiarities of the kinetic behaviour of the free basic group were explained by the hydrophilic nature of the quaternized PVP changing as the degree of quaternization is increased so causing an exclusion of the substrate molecules from the polyion.  相似文献   

3.
The title compound, C36H44N6O4+·2Cl?·2ClO4?·0.132H2O, is shown to be protonated at all the pyridine N atoms; the two chloride ions are hydrogen bonded to three pyridine N atoms and to the phenolic O atom of the same cation [Cl?N = 3.045 (2)–3.131 (2) Å and Cl?O = 2.938 (2) Å], and the remaining pyridine N atom is hydrogen bonded to the phenolic O atom [N?O = 2.861 (2) Å]. The mean value of the C—N—C angle of the protonated pyridine rings is 123.4 (1)°, which is significantly larger than that found for unprotonated pyridine rings.  相似文献   

4.
The interaction between 3-methyl-4-pyrimidone and phenol derivatives or HBr has been studied by IR spectrometry in solution and in the solid state. For pKa values ranging from 10.3 to 4.5, normal OH··· OC hydrogen bonds are formed. With stronger acids (PKa = 0.4 to ?6) protonation occurs at the N(1) nitrogen atom of the ring. For phenols of intermediate pKa values (3.5), there is no preferred site of interaction, both OH···OC and NH+···O? bonds being formed in solution.  相似文献   

5.
Complexes [Sb(2Ac4oClPh)Cl2] (1), [Sb(2Ac4oFPh)Cl2] (2), [Sb(2Ac4oNO2Ph)Cl2] (3), [Sb(2Bz4oClPh)Cl2] (4), [Sb(2Bz4oFPh)Cl2] (5) and [Sb(2Bz4oNO2Ph)Cl2] (6) were obtained with 2-acetylpyridine-N(4)-ortho-chlorophenyl thiosemicarbazone (H2Ac4oClPh) and its N(4)-ortho-fluor (H2Ac4oFPh) and N(4)-ortho-nitro (H2Ac4oNO2Ph) analogues, and with the corresponding 2-benzoylpyridine-derived thiosemicarbazones (H2Bz4oClPh, H2Bz4oFPh, H2Bz4oNO2Ph). The studied compounds are excellent inhibitors of Trypanosoma cruzi growth. H2Bz4oClPh and complexes (4) and (1) were the most trypanosomicidal.Upon coordination of H2Ac4oClPh to antimony(III) in 1, the therapeutic index (TI) goes from 10.58 to 14.35. However, the best values of TI were found for H2Bz4oClPh (TI = 1240) and H2Ac4oNO2Ph (TI = 773). Structure-activity relationship (SAR) studies did not allow the establishment of correlations between the anti-trypanosomal activity and physico-chemical parameters, but correlations were found between the cytotoxicities and physico-chemical properties.  相似文献   

6.
Molybdenum-95 NMR chemical shifts are reported for a series of Mo(O) compounds of the type Mo(CO)5L (L = pyridine derivatives). A good correlation is found between the δ(95Mo) values and the Hammett sigma constant of the pyridine substituent or the pKa of the substituted pyridine. The chemical shift values, which range from −1366 ppm (3-CN, σ = 0.62, pKa = 1.35) to −1433 ppm (4-NMe2, σ = −0.83, pKa = 9.61), directly reflect the electronic properties of the pyridine derivatives even though the substituent is four or five bonds away from the molybdenum atom.  相似文献   

7.
The reaction of alkyl aryl N-p-tosylsulphilimines with thiophenolate ion was found to afford quantitatively the sulphide that arises by an SN2 like reaction on the carbon atom adjacent to the tri-valent sulphur atom. This reaction was also found to proceed smoothly with such compounds as sulphoxides and sulphones and sulphoxmanes. The kinetic study on the reaction between aryl methyl N-p-tosylsulphilimine with thiophenolate ion in DMF reveals that the reaction is of second order, namely, first order with respect to each thiophenolate ion and the sulphilimine. The enthalpy and entropy of activation for the reaction are ΔH = ?17· kcal/mol and ΔS = ?5·7 eu respectively. The effect of substituents in the reaction, p-XC6H4+(?SO2C6H4Y-p)CH3 + p-ZC6H4SK is nicely correl with Hammett σ values giving ?x = + 2·4, ?y = + 1·2 and ?z = ?1·8 respectively. Meanwhile, a marked steric retardation by a bulky alkyl group in alkyl phenyl N-p-tosylsulphilimine is observed. Furthermore, from the stereochemical study of the reaction using an optically active sec-octyl phenyl N-p-tosylsulphilimine with thiophenolate ion it is concluded that the reaction proceeds via a typical SN2 process on α-carbon atom attached to the tri-valent sulphur atom.  相似文献   

8.
The mechanism of hydrolysis of n-nitrophenyl acetate (NPA), butyrate (NPB), caprylate (NPC), and o-methoxycinnamate (NPOMC) catalysed by benzyl-containing polyethyleneimines of linear and branched structures was investigated in aqueous media. The reaction seems to proceed via a general basic mechanism of catalysis and does not involve acylation of the catalyst. Benzyldiethylamine is an analogue of the active centres in polymers with pKa = 8 · 35 ± 0 · 1, localized in the polymer globules at sites of higher hydrophobity.The reaction has a three-step mechanism involving binding of the substrate to an active centre (to give Michaelis sorption complex), substrate conversion and desorption of products. For each step, rate constants were determined. The effect of polymer (K2/Km)/KII increases from NPA to NPC; in the latter case, it is of order 105.  相似文献   

9.
《Tetrahedron: Asymmetry》2001,12(13):1825-1828
Molecular hydrogen is almost four times more soluble in the ionic liquid 1-n-butyl-3-methylimidazolium tetrafluoroborate (BMI·BF4) than in its hexafluorophosphate (BMI·PF6) analogue at the same pressure. The Henry coefficient solubility constant for the solution BMI·BF4/H2 is K=3.0×10−3 mol L−1 atm−1 and 8.8×10−4 mol L−1 atm−1 for BMI·PF6/H2, at room temperature. The asymmetric hydrogenation of (Z)-α-acetamido cinnamic acid and kinetic resolution of (±)-methyl-3-hydroxy-2-methylenebutanoate by (−)-1,2-bis((2R,5R)-2,5-diethylphospholano)benzene(cyclooctadiene)rhodium(I) trifluoromethanesulfonate and dichloro[(S)-(−)-2,2′-bis(di-p-tolylphosphino)-1,1′-binaphthyl]ruthenium(II) complexes immobilised in BMI·PF6 and BMI·BF4 were investigated. Remarkable effects in the conversion and enantioselectivity of these reactions were observed as a function of molecular hydrogen concentration in the ionic phase rather than pressure in the gas phase.  相似文献   

10.
E. Langer  H. Lehner 《Tetrahedron》1973,29(2):375-383
Concerning the question of transanular II-II-interactions in [2.2]metacyclophane, [2.2]paracyclophane and 2,2′-spirobiindane.From the quotient of the two dissociation constants (K1/K2) of [2.2]metacyclophane-bis-chromtricarbonyl (9·0 ± 1·9) it was concluded that there are no transanular II-II-interactions between the two benzene rings. The corresponding values for the bis-chromtricarbonyl-complexes of 2,2′-spirobiindane and [2.2]paracyclophane are 8·0 ± 1·5 and 104, resp. These results are supported by IR-spectroscopical data of the CO-frequencies of the Cr(CO)3-complexes of [2.2]metacyclophane and some derivatives, of 2,2′-spirobiindane and [2.2]paracyclophane.Moreover, UV-spectroscopic studies of tetracyanoethylene complexes of arenes are shown to be insignificant with regard to transanular II-II-interactions.  相似文献   

11.
The formation of VC-SO2 and VC-(SO2)2 complexes in liquid mixtures of vinyl chloride (VC) and sulphur dioxide has been shown by (a) the freezing point composition diagram and (b) chemical shifts in the PMR spectrum of VC over the complete composition range. It is postulated that SO2 can associate with the CC bond and the Cl atom. These complexes may be involved in the copolymerization and influence the composition and stereochemistry of the product. PMR spectra of VC-SO2-ethane(E) mixtures with [SO2] ? [E] ? [VC] gave Kv = 2·0 ± 0·5, 1·5 ± 0·1 and 1·1 ± 0·3 at 232·6, 272·6 and 301·3 K with ΔHf0Hf = ?6·6 ± 1·4 kJ mol?1 for the VC -(SO2)2 complex. The chemical shift of the trans β-proton was twice that of the other two protons. indicating that SO2 adopts an asymmetric orientation to the double bond.  相似文献   

12.
By using the potentiometric titration method, we have determined the pK a values of the two terminal lysine groups in six alanine-based peptides differing in the length of the alanine chain: Ac?CLys?CLys?CNH2 (KK), Ac?CLys?CAla?CLys?CNH2 (KAK), Ac?CLys?CAla?CAla?CLys?CNH2 (KAK2), Ac?CLys?CAla?CAla?CAla?CLys?CNH2 (KAK3), Ac?CLys?CAla?CAla?CAla?CAla?CLys?CNH2 (KAK4), and Ac?CLys?CAla?CAla?CAla?CAla?CAla?CLys?CNH2 (KAK5) in aqueous solution. For each compound, the model of two stepwise acid?Cbase equilibria was fitted to the potentiometric-titration data. As expected, the pK a values of the lysine groups increase with increasing length of the alanine spacer, which means that the influence of the electrostatic field between one charged group on the other decreases with increasing length of the alanine spacer. However, for KAK3, the pK a1 value (8.20) is unusually small and pK a2 (11.41) is remarkably greater than pK a1, suggesting that the two groups are close to each other and, in turn, that a chain-reversal conformation is present for this peptide. Starting with KAK3, the differences between pK a1 and pK a2 decrease; however, for the longest peptide (KAK5), the values of pK a1 and pK a2 still differ by about 1 unit, i.e., by more than the value of log10 (4)?=?0.60 that is a limiting value for the pK a difference of dicarboxylic acids with increasing methylene-spacer length. Consequently, some interactions between the two charged groups are present and, in turn, a bent shape occurs even for the longest of the peptides studied.  相似文献   

13.
The reactivity of the coordinated monodentate oxine and of the bidentate anion in the complexes ThOx4·HOx and UO2Ox2·HOx (HOx = oxine) were compared by the reaction of these complexes with (a) haloacetic acids and (b) nitrophenols. The mixed ligand complexes isolated from (a) were formulated as ThOxy (RCO2)x·nHOx, for R = CCl3, y = 2, x = 2, n = 0 and for R = CHCl2, y = 1, x = 3, n = 1, and as UO2Oxy(RCO2)x·nHOx, for R = CF3 or CH2CL, y = x = n = 1, and for R = CCl3, y = 0, x = n = 2. In the case of (b) the formed molecular complexes were formulated as ThOx4·2 Hph and as UO2Ox2·Hph [Hph = picric acid and 2,4-dinitrophenol for Th(IV) and also 2,6-dinitrophenol for U(VI)]. A hydrogen bonding formation that combines the oxygen atom of the chelated oxinate and that of the nitrophenol is suggested. These tentatively suggested structures are in accordance with analytical and spectral evidence.  相似文献   

14.
Pentaquadrupole (QqQqQ) mass spectrometry is used to explore the abilities of gaseous SFn+ (n = 1–5) ions to form adducts and dimers with three π-electron rich molecules—benzene, acetonitrile, and pyridine, whereas ab initio calculations estimate most feasible structures, bond dissociation energies (BDEs), and reaction enthalpies of the observed products. With benzene, SF+ reacts by net H-by-SF replacement. As suggested by the calculations, this novel benzene reaction forms ionized benzenesulfenyl fluoride, C6H5–SF, via a Wheland-type intermediate that spontaneously loses a H atom. SF3+ forms a rare, loosely bonded π complex with benzene, [Bz ⋯ SF3]+, which is stable toward both H and HF loss. No dimer, Bz2SF3+, is formed. According to calculations, an unsymmetrically bonded, π-coordinated Bz2SF3+ dimer exists, i.e. (Bz–SF3 ⋯ Bz)+, but its formation from [Bz ⋯ SF3]+ is endothermic; hence, thermodynamically unfavorable. With acetonitrile, SF2, SF3+, and SF5+ form both adducts and dimers. CH3–C·N–SF2+ (a new distonic ion) and CH3CN–SF5+ are covalently bonded, but CH3CN ⋯ SF3+ is loosely bonded. The binding natures of the acetonitrile adducts are reflected in the dimers; [CH3CN–SF2 ⋯ NCCH3] and [CH3CN–SF5 ⋯ NCCH3]+ are unsymmetrically bonded, whereas [CH3CN ⋯ SF3 ⋯ NCCH3]+ is symmetrically and loosely bonded. Such dimers as [CH3CN ⋯ SF3 ⋯ NCCH3]+ are ideal for measurements of ion affinity via the Cooks’ kinetic method. With pyridine, only SF3+ forms adduct and dimer. Py–SF3+ is covalently bonded through nitrogen; [Py ⋯ SF3 ⋯ Py]+ is loosely but unsymmetrically bonded. The unsymmetric 2.28 and 2.44 Å long N–S bonds in [Py ⋯ SF3 ⋯ Py]+, which are expected to rapidly interconvert, result likely from steric hindrance that forces orthogonal alignment of the two pyridine rings. Most observed adducts and dimers display relatively high BDEs, i.e. they are formed in thermodynamically favorable reactions. The extents of dissociation of the adducts and dimers observed in MS3 experiments reflect the structures and BDEs predicted by the calculations.  相似文献   

15.
L.S. Levitt  B.W. Levitt 《Tetrahedron》1975,31(19):2355-2357
The first ionization potentials, EI, of substituted chromium(III) acetylacetonates are found to be a linear function of the sum of Hammett's σm and σp constants for substituents in various positions of the pseudo-aromatic rings, relative to the Cr atom. The equation for the regression is EI = 7·40 + 0·380Σσ ± 0·02 eV, where 7·40 eV is the interpolated value for the tris-Cr chelate of malonaldehyde, and the slope of the correlation line is 0·380. From the value of the slope and other considerations, it is deduced that the ejected electron is an unpaired dxyCr electron in at2gMO.  相似文献   

16.
The ionic complexes simultaneously containing negatively charged coordination structures of metal phthalocyanines and fullerene anions, viz., {MnIIPc(CH3CH2S?) x ·(I?)1?x }·(C60 ·?)· ·(PMDAE+)2·C6H4Cl2 (PMDAE is N,N,N′,N′,N′-pentamethyldiaminoethane, x = 0.87, 1) and {ZnIIPc(CH3CH2S?)y·(I?)1?y }2·(C60 ?)2·(PMDAE+)4·(C6H4Cl2) (y = 0.5, 2) were synthesized. The both compounds were obtained as single crystals, which made it possible to study their crystal structures. In complex 1, the fullerene radical anions form honeycomb-like layers in which each fullerene has three neighbors with center-to-center interfullerene distances of 10.13–10.29 Å. Rather long distances between the C60 ·? radical anions results in the retention of monomeric C60 ·? in this complex down to the temperature of 110(2) K. In complex 2, fullerenes form dimers (C60 ?)2 bonded by one C-C bond. The dimers are packed in corrugated honeycomb-like layers with interfullerene center-to-center distances of 9.90–10.11 Å. Manganese(II) and zinc(II) phthalocyanines coordinate iodide and ethanethiolate anions to the central metal atom to form unusual negatively charged coordination structures MIIPc(An?) (An? is anion) packed in dimers {MIIPc(An?)}2 with a short distance between the phthalocyanine planes (3.14 Å in 1 and 3.27 Å in 2). The pthalocyanine dimers also form layers with the PMDAE+ cations, and these layers alternate with the fullerene layers. The packing of spherical fullerenes with planar phthalocyanine molecules is attained by the insertion of fullerenes between the phenylene groups of phthalocyanines. The π-π-interactions of the porphyrin macrocycle with five- or six-membered fullerene rings are characteristic of the earlier studied ionic porphyrin and fullerene complexes. Such interactions are not observed for ionic complexes 1 and 2.  相似文献   

17.
Condensation of di-2-pyridyl ketone with S-methyldithiocarbazate or S-benzyldithiocarbazate yields potentially bridging ligands of the form Py2CNNHC(S)SR; Hdpksme (R = Me; the di-2-pyridyl ketone Schiff base of S-methyldithiocarbazate) and Hdpksbz (R = Bz; the di-2-pyridyl ketone Schiff base of S-benzyldithiocarbazate). Complexation of these ligands with Cu(II) in a 1:1 M ratio leads to the formation of dinuclear complexes of the general formula [Cu(NNNS)X]2 (X = Cl, NO3, H2O). X-ray crystallographic structure determinations show that each ligand provides three donor atoms (NNS) in a meridional configuration to one metal, viz. one of the pyridine nitrogen atoms, the azomethine nitrogen atom and the thiolate sulfur, while the nitrogen atom of the second pyridyl group forms a bridge to another copper(II) ion within the dimer. The coordination geometry around each copper(II) ion is approximately square pyramidal, the basal plane of which is composed of one of the pyridine nitrogen atoms, the azomethine nitrogen atom and a chlorido, nitrato or aqua ligand. The apical position of the square pyramid is always occupied by the pyridine nitrogen atom of the second ligand.  相似文献   

18.
The highly insoluble organic-inorganic hybrid ionic compounds N,N??-methylenedipyridinium tetrachloroplatinate(II) [(C5H5N)2CH2] · [PtCl4] and N,N??-methylenedipyridinium hexachloroplatinate(IV) [(C5H5N)2CH2] · [PtCl6] were obtained by the treatment of N,N??-methylenedipyridinium dichloride monohydrate [(C5H5N)2CH2]Cl2 · H2O with K2[PtCl4] or (NH4)2[PtCl6], respectively, in an aqueous solution. Both complexes were isolated, purified, characterised by elemental analysis, and their molecular structures were confirmed by powder X-ray diffraction. The crystal structure of both compounds consists of separated discrete dications [(C5H5N)2CH2]2+ and anions [PtCl n ]2? (n = 4 or 6). As anticipated, the dications formed a butterfly shape consisting of two pyridine rings bound to the methylene group via their N atoms, while the Pt centre had a square planar geometry in [(C5H5N)2CH2] · [PtCl4] and an octahedral coordination in [(C5H5N)2CH2] · [PtCl6]. Interestingly, both crystal structures are stabilised by intermolecular C-H??Cl non-standard hydrogen bonds, ??-?? ring interactions between two pyridine rings of adjacent dications, and also by Cl-?? interactions.  相似文献   

19.
The protonation constants and solubilities of three complexons [ethylenediamine-N,N′-disuccinic acid (EDDS), ethylene glycol bis(2-aminoethyl ether)-N,N,N′,N′-tetraacetic acid (EGTA) and 1,2-cyclohexanediamine-N,N,N′,N′-tetraacetic acid (CDTA)] are reported in aqueous solutions of NaCl with different ionic strength values (0 ≤ I ≤ 4.8 mol·L?1) and, in the case of CDTA, in (CH3)4NCl (0.1 ≤ I ≤ 2.7 mol·L?1). The dependence on ionic strength of the protonation constants of these three complexons and four other complexons that were previously reported (NTA, EDTA, DTPA and TTHA), is analyzed in NaCl solution; the ionic strength influences quite strongly the protonation constants (as an example for CDTA, log10 K 1 = 10.54 and 9.25 at I = 0.1 and 1 mol·L?1, respectively), while the effect of (CH3)4NCl concentration is lower. Based on the total solubility S T and the protonation constant data at different salt concentrations, the solubility of the neutral species S 0 and the solubility products K S0 are obtained. The Setschenow coefficients k m and the solubility values S 0 0 in pure water are also reported (S 0 0  = 0.55, 0.21 and 0.75 mmol·kg?1 for EDDS, EGTA and CDTA, respectively). The dependence of the protonation constants on ionic strength is also interpreted in terms of ion pair formation, and the formation constants of Na+ species are reported.  相似文献   

20.
Five Mn(II) complexes of bis(thiosemicarbazones) which are represented as [Mn(H2Ac4Ph)Cl2] (1), [Mn(Ac4Ph)H2O] (2), [Mn(H2Ac4Cy)Cl2]·H2O (3), [Mn(H2Ac4Et)Cl2]·3H2O (4) and [Mn(H2Ac4Et)(OAc)2]·3H2O (5) have been synthesized and characterized by elemental analyses, electronic, infrared and EPR spectral techniques. In all the complexes except [Mn(Ac4Ph)H2O], the ligands act as pentadentate neutral molecules and coordinate to Mn(II) ion through two thione sulfur atoms, two azomethine nitrogens and the pyridine nitrogen, suggesting a heptacoordination. While in compound [Mn(Ac4Ph)H2O], the dianionic ligand is coordinated to the metal suggesting six coordination in this case. Magnetic studies indicate the high spin state of Mn(II). Conductivity measurements reveal their non-electrolyte nature. EPR studies indicate five g values for [Mn(Ac4Ph)H2O] showing zero field splitting.  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