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.
Large, inert, weakly basic carborane anions of the icosahedral type CHB(11)R(5)X(6)(-) (R = H, Me; X = Cl, Br) allow ready isolation and structural characterization of discrete salts of the solvated proton, [H(solvent)(x)][CHB(11)R(5)X(6)], (solvent = common O-atom donor). These oxonium ion Br?nsted acids are convenient reagents for the tuned delivery of protons to organic solvents with a specified number of donor solvent molecules and with acidities leveled to those of the chosen donor solvent. They have greater thermal stability than the popular [H(OEt(2))(2)][BAr(F)] acids based on fluorinated tetraphenylborate counterions because carborane anions can sustain much higher levels of acidity. When organic O-atom donors such as diethyl ether, tetrahydrofuran, benzophenone, and nitrobenzene are involved, the coordination number of the proton (x) in [H(solvent)(x)()](+) is two. A mixed species involving the [H(H(2)O)(diethyl ether)](+) ion has also been isolated. These solid-state structures provide expectations for the predominant molecular structures of solvated protons in solution and take into account that water is an inevitable impurity in organic solvents. The O.O distances are all short, lying within the range from 2.35 to 2.48 A. They are consistent with strong, linear O.H.O hydrogen bonding. Density functional theory calculations indicate that all H(solvent)(2)(+) cations have low barriers to movement of the proton within an interval along the O.H.O trajectory, i.e., they are examples of so-called SSLB H-bonds (short, strong, low-barrier). Unusually broadened IR bands, diagnostic of SSLB H-bonds, are observed in these H(solvent)(2)(+) cations.  相似文献   

2.
Solutions of benzenesulfonic acid (BSA) and 4-toluenesulfonic acid monohydrate (PTSA) in dimethylsulfoxide (DMSO) were studied by FT-Raman spectroscopy in the concentration range 1.0-3.5 mol dm(-3) (BSA) and 1.0-4.8 mol dm(-3) (PTSA). Spectra in the region of the Raman acid complex band (C-S + C-C + SO3) stretches, at 1124 cm(-1) were analysed by band-fitting procedures in order to ascertain the degree of acid dissociation. In BSA solutions, this parameter changes from 0.78 at 1.02 M to 0.47 at 3.5M, despite the strong character of the acid. Interaction of DMSO with undissociated BSA produces a new band in the solvent nu(C-S) Raman spectral region near 671 cm(-1), displaced >15.0 cm(-1), and assigned to DMSO molecules H-bonded to BSA. In PTSA solutions, hydrogen bonds are formed with the oxonium ion (H3O+) dissociated from the acid. In this case, the displacement observed is only >10.0 cm(-1), indicating a weaker interaction. From the concentration of H-bonded DMSO, the solute/solvent coordination number and its inverse, the mean number of H-bonds participating in bonding with each solvent molecule can be calculated. This coordination number changes in BSA solutions in bimodal way, passing through a maximum and reaching a limit of 2 in the most concentrated solution. This number agrees with that found in the solid solvate BSA.2DMSO. In PTSA solutions, the general trend is similar, but low coordination numbers are obtained, in agreement with the low acidity of the oxonium ion. The bimodal behaviour observed in both acids is explained by the self-associated structure of the solvent.  相似文献   

3.
Partition coefficients P of the HNCS, HNCO and HN3 hydropseudohalic acids between a number of organic solvents and water were determined. It has been found that log P increases with pKa of the acid and with the basicity of the solvent, but the effect of pKa on P is the smaller the more basic is the solvent. The relationships have been explained in terms of hydrogen bond formation between undissociated acid and solvent molecules. H-bonding between the pseudohalic acids and organic solvents has been confirmed by IR spectra on the example of HN3 in benzene. Association constants for H-bonding between the three acids and water, benzene, dibutyl ether and tri-n-butyl phosphate were determined from partition data. It has been found that H-bonding increases with the strength of the acid, whereas the contribution to partition from non-specific interactions with water and organic solvents depends on the molecular surface area of the acid molecule.  相似文献   

4.
Two novel compounds, (L(1)H)(2)[SiF(6)] x 2H(2)O (1) and (L(2)H)(2)[SiF(5)(H(2)O)](2) x 3H(2)O (2), resulting from the reactions of H(2)SiF(6) with 4'-aminobenzo-12-crown-4 (L(1)) and monoaza-12-crown-4 (L(2)), respectively, were studied by X-ray diffraction and characterised by IR and (19)F NMR spectroscopic methods. Both complexes have ionic structures due to the proton transfer from the fluorosilicic acid to the primary amine group in L(1) and secondary amine group incorporated into the macrocycle L(2). The structure of 1 is composed of [SiF(6)](2-) centrosymmetric anions, N-protonated cations (L(1)H)(+), and two water molecules, all components being bound in the layer through a system of NH[...]F, NH[...]O and OH[...]F hydrogen bonds. The [SiF(6)](2-) anions and water molecules are assembled into inorganic negatively-charged layers via OH[dot dot dot]F hydrogen bonds. The structure of 2 is a rare example of stabilisation of the complex anion [SiF(5)(H(2)O)](-), the labile product of hydrolytic transformations of the [SiF(6)](2-) anion in an aqueous solution. The components of 2, i.e., [SiF(5)(H(2)O)](-), (L(2)H)(+), and water molecules, are linked by a system of NH[...]F, NH[...]O, OH[...]F, OH[dot dot dot]O hydrogen bonds. In a way similar to 1, the [SiF(5)(H(2)O)](-) anions and water molecules in 2 are combined into an inorganic negatively-charged layer through OH[...]F and OH[...]O interactions.  相似文献   

5.
It is demonstrated that dissolution of aminothiacalix[4]arene in chloroform results in transformation of 1,3-alternate conformation, adopted in single-crystal and bulk polycrystalline solids, to the pinched-cone form. This conformer is stabilised by the intramolecular hydrogen bonds of two distal amino-groups acting as H-donors with another two amino moieties that appear as H-acceptors. The H-bonds cause quite small (ca. 10–20 cm?1) red shift of the IR bands of the NH2 stretching vibrations, which suggests rather weak NH?N hydrogen bonding. This latter is sufficient to stabilize the pinched-cone conformation in the chloroform solution, but the energy gap between the pinched-cone and other conformations is small, and solid-state intermolecular forces easily overcome it, leading to realisation of the 1,3-alternate conformer. The comparison of the DFT computed and experimental vibrational and NMR spectra demonstrates good quality of present quantum-chemical computations, allows complete interpretation of the spectra and reveals simple IR and NMR spectroscopic markers of the conformers of aminothiacalix[4]arenes.  相似文献   

6.
Carboxyl (COO?) vibrational modes of two amino acids histidine and glycine in D2O solution were investigated by temperature-dependent FTIR spectroscopy and temperature-jump nanosecond time-resolved IR di?erence absorbance spectroscopy. The results show that hydrogen bonds are formed between amino acid molecules as well as between the amino acid molecule and the solvent molecules. The asymmetric vibrational frequency of COO? around 1600-1610 cm?1 is blue shifted when raising temperature, indicating that the strength of the hydrogen bonds becomes weaker at higher temperature. Two bleaching peaks at 1604 and 1612 cm?1 were observed for histidine in response to a temperature jump from 10 ±C to 20 ±C. The lower vibrational frequency at 1604 cm?1 is assigned to the chain COO? group which forms the intermolecular hydrogen bond with NH3+ group, while the higher frequency at 1612 cm?1 is assigned to the end COO? group forming hydrogen bonds with the solvent molecules. This is because that the hydrogen bonds in the former are expected to be stronger than the latter. In addition the intensities of these two bleaching peaks are almost the same. In contrast, only the lower frequency at 1604 cm?1 bleaching peak has been observed for glycine. The fact indicates that histidine molecules form a dimer-like intermolecular chain while glycine forms a relatively longer chain in the solution. The rising phase of the IR absorption kinetics in response to the temperature-jump detected at 1604 cm?1 for histidine is about 30§10 ns, within the resolution limit ofour instrument, indicating that breaking or weakening the hydrogen bond is a very fast process.  相似文献   

7.
The alkoxides TpPh,MeZn-OR (R = Me, Et, i-Pr) undergo acid-base reactions with all hydrogen compounds whose acidity is higher than that of the corresponding alcohol ROH. Thus, anion exchange occurs with the common acids acetic acid, acetohydroxamic acid, acetylacetone, phenol, and ethylmercaptan. Alkoxide exchange is observed using methanol, ethanol, and trifluoroethanol. With the NH acids cyanamide, trifluoroacetamide, and pyrazoles, the corresponding anions are attached to zinc, and likewise beta- and gamma-lactams, a thiazolidinedione, and the cyclic sulfimide saccharin are deprotonated. Of the CH acids acetonitrile forms the Tp*Zn-cyanomethanide. Acetone is deprotonated by the cyanomethanide complex and incorporated as the Tp*Zn-beta-ketoiminate.  相似文献   

8.
壳聚糖-两性聚氨酯接枝共聚物的制备   总被引:2,自引:0,他引:2  
利用壳聚糖大分子链上的-OH和-NH2基团与两性聚氨酯预聚体的端-NCO基团反应,把两性聚氨酯链接枝到壳聚糖分子链上,为聚电解质及纳米胶体粒子间的组装,复合性质的研究提供模型化合物,红外光谱,热分析方法初步研究表明,采用上述方法能够制备出APU-壳聚糖接枝共聚物,但产物是壳聚糖与接枝共聚物的混合物,研究发现,这种接枝共聚物在盐酸水溶液中呈胶束形态,而且具有独特的流变性。  相似文献   

9.
Fundamental model compounds for the Hammett equation, meta- and para-substituted benzoic acids, were investigated by the density functional theory at the B3LYP/6-311+G(d,p) level. Energies of 25 acids and of their anions were calculated in all possible conformations and from them the energies of the assumed mixture of conformers. Relative acidities correlated with the experimental gas-phase acidities almost within the experimental uncertainty, much more precisely than in the case of previous calculations at lower levels. Dissection of the substituent effects into those operating in the acid molecule and in the anion was carried out by means of isodesmic reactions starting from monosubstituted benzenes. Both effects are cooperating in the resulting effect on the acidity; those in the acid molecule are smaller but not negligible. They are also responsible for some deviations from the Hammett equation (through-resonance of para donor substituents) and for the weaker resonance in the acid molecule in meta derivatives; in the anions the inductive and resonance effects are almost equal. On the other hand, the cooperation of effects in the acid and in the anion makes the relative acidity more sensitive to electron withdrawing and is probably one of the reasons why the Hammett equation is so generally valid.  相似文献   

10.
Density functional theory (DFT) calculations have been performed to investigate the gas-phase conformations of serine and its three related ions (serineH(+), serine(-), and serine(2-)). The full ensemble of possible conformations, 324 conformations for serine, 108 for serineH(+), 162 for serine(-) and 54 for serine(2-), were first surveyed at B3LYP/6-31G level, and then the obtained unique conformations were further refined at B3LYP/6-311+G level. From full optimizations, 74 unique conformations for seine, 14 for serineH(+), 11 for serine(-), and 4 for serine(2-) were located, and their relative energies were also determined at B3LYP/6-311+G level. Atoms in molecules (AIM) analysis was carried out to establish rigorous definition of hydrogen bonds. Six types of intramolecular H-bonds in conformers of serine, six types in serineH(+), three types in serine(-), and two types in serine(2-) were identified within the framework of AIM theory and their relative strengths were determined based on topological properties at bond critical points (BCPs) of H-bonds. The intramolecular H-bonds were demonstrated to play an important role in deciding the relative stability of conformations of amino acids and the related ions. The enthalpies and Gibbs free energies of protonation and deprotonation reactions of serine and its related ions were calculated at B3LYP/6-311+G//B3LYP/6-31G, and B3LYP/6-311+G//B3LYP/6-311+G level. The calculated results are both in excellent agreement with the experimental data. We demonstrate in this study that B3LYP is an efficient and accurate method to predict the thermochemical and structural parameters of amino acids and the related ions.  相似文献   

11.
Intra- and intermolecular interactions and the conformations of 3-(2-bromo-pyridinium)-propionic acid bromide (1) and 3-(pyridinium)-pivalic acid chloride (2) have been studied by X-ray diffraction and theoretically. In these complexes, Br and Cl anions are H-bonded to the COOH group and interact electrostatically with the positively charged nitrogen atoms of neighbouring molecules. To analyse these interactions theoretically, the structures of monomers and dimers in various configurations were optimized by PM3 and BLYP/6-31G(d,p) methods. The analysis confirmed the vital role of both the H-bonds and electrostatic interactions (intra- and intermolecular) for the molecular conformation and the ionic aggregation in the solid state.  相似文献   

12.
The ionization and solution enthalpies of the nitrophenol isomers were measured calorimetrically at 25°C in water-DMSO mixtures ranging from pure water up to 0.8 DMSO mole fraction.The different solvation values obtained for ortho-nitrophenol with respect to the meta and para isomers, were explained taking into account the electrostatic forces between the water molecules and the anions, as well as the intramolecular and intermolecular hydrogen bonds associated with the undissociated molecules.  相似文献   

13.
The relatively strong acidity of hydroxamic acids was analyzed by means of isodesmic reactions in which this acid or its anion is formed from simpler precursors. Acidity of amides was analyzed in the same way. Energies of all compounds involved in the reactions were calculated at the B3LYP/AUG-cc-pVTZ//B3LYP/6-311 + G(d,p) level; at this level a good agreement was reached with the sparse experimental data. Interpretation of the results was the same as in the recent discussion of the acidity of carboxylic acids, and the conclusions were similar: both amides and hydroxamic acids are stabilized with respect to simpler reference molecules of amines or N-alkylhydroxylamines, respectively. However, their anions are stabilized still more and are responsible for the acidity. This effect is stronger in hydroxamic acids or amides than in carboxylic acids. The problem of whether it is due to resonance depends on the definition of this term. Semiquantitative comparison suggests that resonance in hydroxamic acids is more important than in amides and still more than in carboxylic acids. The stronger acidity of hydroxamic acids compared to amides is due to the destabilizing inductive effect of the hydroxyl group in the acid molecule, not to any effect in the anion.  相似文献   

14.
The ionization and solution enthalpies of the hydroxybenzoic acids were measured calorimetrically at 25°C in water—dimethylsulfoxide mixtures ranging from pure water to 0.8 DMSO mole fraction.In the same solution, the ΔG° values for the ionization processes have been determined by potentiometric measurements.The different enthalpy solvation values for the three hydroxy isomers were explained taking into account the prevalence of the resonance or the inductive effect as a function of the solvent composition.The intramolecular hydrogen bond of the undissociated molecules seems to be responsible for the greater degree of ionization of o-hydroxybenzoic acid with respect to the m and p-hydroxy isomers.  相似文献   

15.
The structure of trifluoro-N-(3-formylcyclohept-2-en-1-yl)methanesulfonamide and its self-association in solution have been studied by IR spectroscopy and quantum chemical methods [B3LYP/6-311G(d,p), AIM]; proton affinities of basic centers in its molecule have been evaluated. Trifluoro-N-(3-formylcyclohept-2-en-1-yl)methanesulfonamide in inert solvents forms cyclic dimers, whereas in crystal chain associates are more likely to be formed via hydrogen bonding between the NH and C=O groups of neighboring molecules. The carbonyl group in the title compound undergoes protonation only by the action of very strong trifluoromethanesulfonic acid. Weaker acids give rise to solvate H-complexes at the NH, C=O, and S=O groups. The topology of hydrogen bonds in dimers of different types has been analyzed in terms of the AIM theory.  相似文献   

16.
Di- and trinitrophenide anions generated by decarboxylation of the anions of 2,4-, 3,5-, and 2,6-dinitrobenzoic acids and 1,3,5-trinitrobenzoic acid in the medium-pressure region of an electrospray ion source react locally with various C-H acids delivered in the form of vapors mixed with the curtain gas, yielding anionic sigma-adducts. Positive results were obtained for aliphatic aldehydes, ketones, esters and nitriles. All three dinitrobenzoic acids bearing NO(2) groups in the meta position to each other gave the same sigma-adducts which can be rationalized by a reaction sequence including proton transfer from the C-H acid to the nitrophenide anion and subsequent formation of the sigma-adduct by the reaction of 1,3-dinitrobenzene with the carbanion within the ion-molecule complex. It was found that such a reaction is possible only for C-H acids with a gas-phase acidity lying within a narrow, strictly defined range whose location on the acidity scale depends on the acidity of the nitroarene. The sigma-adduct formed in the reaction of the 2,4-dinitrophenide anion with CH(2)Cl(2) undergoes rapid HCl elimination yielding an anion with the same composition as that produced by the Vicarious Nucleophilic Substitution of hydrogen reaction but with a different structure.  相似文献   

17.
The Raman and IR spectra of NH3(CH2)5NH3SnCl6 have been measured at ambient temperature. It is shown that the cations in the compound assume a symmetry lower than C2v. Combination bands observed in the 2100-1800 cm(-1) region in the IR spectrum of NH3(CH2)5NH3SnCl6 indicate that the compound contains the C-NH3 grouping, the bands are discussed and their assignment are suggested. No evidence of existence of hydrogen bonding is found from the infrared spectrum in the region of 2800-3200 cm(-1); anions and cations are found not connected by hydrogen bonding and are therfore isolated. The Raman spectrum of anions can be interpreted in terms of disordered groups, not clearly showing the predicted splitting of bands.  相似文献   

18.
We have investigated, using NMR, IR, and CD spectroscopy and X-ray crystallography, the conformational properties of peptides 1-10 of beta(3)-aminoxy acids (NH(2)OCHRCH(2)COOH) having different side chains on the beta carbon atom (e.g., R = Me, Et, COOBn, CH(2)CH(2)CH=CH(2), i-Bu, i-Pr). The beta N-O turns and beta N-O helices that involve a nine-membered-ring intramolecular hydrogen bond between NH(i)(+2) and CO(i), which have been found previously in peptides of beta(2,2)-aminoxy acids (NH(2)OCH(2)CMe(2)COOH), are also present in those beta(3)-aminoxy peptides. X-ray crystal structures and NMR spectral analysis reveal that, in the beta N-O turns and beta N-O helices induced by beta(3)-aminoxy acids, the N-O bond could be either anti or gauche to the C(alpha)-C(beta) bond depending on the size of the side chain; in contrast, only the anti conformation was found in beta(2,2)-aminoxy peptides. Both diamide 1 and triamide 9 exist in different conformations in solution and in the solid state: parallel sheet structures in the solid state and predominantly beta N-O turn and beta N-O helix conformations in nonpolar solvents. Theoretical studies on a series of model diamides rationalize very well the experimentally observed conformational features of these beta(3)-aminoxy peptides.  相似文献   

19.
Mixtures of derivatives of Hanabusa's bolaamphiphilic amidocarbamates containing two Z-valinyl subunits and aliphatic spacers that range from ethylenic to octamethylenic are able to form organogels. A coassembly of them is observed in their acetonitrile and toluene gels; namely, the concentration of a given compound at which a gel is formed is lowered by the presence of equimolar quantities of any other compound in the series. The aggregates were studied by wide-angle X-ray diffraction (WAXD) and the results can be rationalized if the gel fibers are formed by supramolecular copolymers. NMR studies reveal that in solution these molecules adopt folded conformations containing intramolecular H-bonds, but IR studies indicate that these are not present in their aggregates. Additionally, analogues of the amidocarbamates obtained by replacement of the carbamate functionality by amide have been shown to behave in a similar way. For these molecules it can be shown that the central aliphatic subunit is not completely extended in the conformations present in the aggregates according to IR and WAXD studies. The tetraamide-type compounds described are robust organogelators that form gels in a variety of organic solvents with good thermostability and present improved feasibility for the synthesis of envisaged functional organogelators.  相似文献   

20.
In continuation of the work on establishing hydrogen donors in the hydrogen evolution reaction from different acid molecules, cathodic evolution of hydrogen on silver is investigated from solutions of monobasic acetic acid with the aim to establish the origin of reduced hydrogen. Solutions of 0.2 M acetic acid with 0.2 M perchloric acid, neutralized to different pH values by NaOH, are used. The earlier established criterion is used for discerning between two possible hydrogen evolution mechanisms: (1) from dissociated hydrogen ions and (2) from undissociated hydrogen atoms in the molecule. At medium pH values, the undissociated acid molecules participate as hydrogen donors. Rate constants for reactions 1 and 2, evaluated at a potential of –800 mV (SCE), at which the entire pH range can be scanned, are 2.9 × 10–6 and 1.9 × 10–8.  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