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 306 毫秒
1.
Acrylic anhydride (AAn) and methacrylic anhydride (MAAn) were polymerized with radical initiator in polar solvents at high temperatures. The polymers obtained were found to consist of five-and six-membered ring structures by comparing IR spectra of the polymers with those of model compounds, succinic anhydride, and glutaric anhydride. Hydrolysis and methylation of the polymers gave new polymers composed of head-to-head (HH) and head-to-tail (HT) methyl acrylate (MA) or methyl methacrylate (MMA) units. The content of HH unit of these HH/HT polymers was determined by 1H-NMR and 13C-NMR spectra. The softening, glass transition, and thermal degradation temperatures of the poly(MA) with HH and HT units were found to somewhat increase with increasing of the content of the HH units. On the other hand, the glass transition and thermal degradation temperatures of the poly(MMA) with HH and HT units increased similarly, but the softening temperature decreased as the content of the HH units increased.  相似文献   

2.
The pharmacokinetic profiles of the hypoglycemic agent, acetohexamide (AH) and its major active metabolite, hydroxyhexamide (HH) were studied in three species of laboratory animals after intraperitoneal (ipl) administration in comparison with those after intravenous (iv) administration of AH and of the preformed metabolite HH. Reductive biotransformation of AH to HH was reversible in rats and guinea pigs, while it was irreversible in rabbits. The parameters of reversible drug-metabolite pharmacokinetics were calculated, including essential clearances of reversible and irreversible elimination, volumes of distribution at the steady state and sojourn times or turnover rates of the metabolite pair. An interconversion model, which incorporated a first-pass metabolism, was applied to the disposition kinetics of AH and HH, and the available fractions of AH and generated metabolite HH in each species were elucidated.  相似文献   

3.
Large reductions in the thermal conductivity of thermoelectrics using nanostructures have been widely demonstrated. Some enhancements in the thermopower through nanostructuring have also been reported. However, these improvements are generally offset by large drops in the electrical conductivity due to a drastic reduction in the mobility. Here, we show that large enhancements in the thermopower and electrical conductivity of half-Heusler (HH) phases can be achieved simultaneously at high temperatures through coherent insertion of nanometer scale full-Heusler (FH) inclusions within the matrix. The enhancements in the thermopower of the HH/FH nanocomposites arise from drastic reductions in the "effective" carrier concentration around 300 K. Surprisingly, the mobility increases drastically, which compensates for the decrease in the carrier concentration and minimizes the drop in the electrical conductivity. Interestingly, the carrier concentration in HH/FH nanocomposites increases rapidly with temperature, matching that of the HH matrix at high temperatures, whereas the temperature dependence of the mobility significantly deviates from the typical T(-α) law and slowly decreases (linearly) with rising temperature. This remarkable interplay between the temperature dependence of the carrier concentration and mobility in the nanocomposites results in large increases in the power factor at 775 K. In addition, the embedded FH nanostructures also induce moderate reductions in the thermal conductivity leading to drastic increases in the ZT of HH(1 - x)/FH(x) nanocomposites at 775 K. By combining transmission electron microscopy and charge transport data, we propose a possible charge carrier scattering mechanism at the HH/FH interfaces leading to the observed anomalous electronic transport in the synthesized HH(1 - x)/FH(x) nanocomposites.  相似文献   

4.
Neopentane and TMS are used as model M(CH(3))(4) systems to investigate intramolecular interactions. The nonbonded site-site potential between two proximal hydrogen atoms on different methyl groups, V(nb)(d(HH)), is not Lennard-Jones- or Morse-like but is found to be pseudolinear in hydrogen-hydrogen internuclear separation, d(HH), for both neopentane and TMS. The Morse potential is found to be a poor basis in which to expand V(nb)(d(HH)). The nonbonded site-site potential is conformation-dependent and not transferable between molecules. The individual contributions to V(nb)(d(HH)) are presented. The local mode parameters for neopentane and TMS are calculated ab initio for a variety of molecular conformations. The ab initio values of the local mode frequency and local mode anharmonicity are increasingly blue-shifted with increasing steric hindrance. Electron correlation is found to be increasingly important with decreasing internuclear separations, d(HH).  相似文献   

5.
Bulk Zr0.25Hf075NiSn half-Heusler (HH) nanocomposites containing various mole fractions of full-Heusler (FH) inclusions were prepared by solid state reaction of pre-synthesized HH alloy with elemental Ni at 1073 K. The microstructures of spark plasma sintered specimens of the HH/FH nanocomposites were investigated using transmission electron microscopy and their thermoelectric properties were measured from 300 K to 775 K. The formation of coherent FH inclusions into the HH matrix arises from solid-state Ni diffusion into vacant sites of the HH structure. HH(1–y)/FH(y) composites with mole fraction of FH inclusions below the percolation threshold, y∼0.2, show increased electrical conductivity, reduced Seebeck coefficient and increased total thermal conductivity arising from gradual increase in the carrier concentration for composites. A drastic reduction (∼55%) in κl was observed for the composite with y=0.6 and is attributed to enhanced phonon scattering due to mass fluctuations between FH and HH, and high density of HH/FH interfaces.  相似文献   

6.
The amplitude and the phase of cross peaks in conventional 2D HSQC experiments are modulated by both proton–proton, J(HH), and proton–carbon, 1J(CH), coupling constants. It is shown by spectral simulation and experimentally that J(HH) interferences are suppressed in a novel perfect‐HSQC pulse scheme that incorporates perfect‐echo INEPT periods. The improved 2D spectra afford pure in‐phase cross peaks with respect to 1J(CH) and J(HH), irrespective of the experiment delay optimization. In addition, peak volumes are not attenuated by the influence of J(HH), rendering practical issues such as phase correction, multiplet analysis, and signal integration more appropriate. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

7.
Experimental and theoretical methods have been used to correlate (2)J(HH) and (3)J(HH) values within the exocyclic hydroxymethyl groups (CH(2)OH) of saccharides with specific molecular parameters, and new equations are proposed to assist in the structural interpretation of these couplings. (3)J(HH) depends mainly on the C-C torsion angle (omega) as expected, and new Karplus equations derived from J-couplings computed from density functional theory (DFT) in a model aldopyranosyl ring are in excellent agreement with experimental values and with couplings predicted from a previously reported general Karplus equation. These results confirm the reliability of DFT-calculated (1)H-(1)H couplings in saccharides. (2)J(HH) values depend on both the C-C (omega) and C-O (theta) torsions. Knowledge of the former, which may be derived from other parameters (e.g., (3)J(HH)), allows theta to be evaluated indirectly from (2)J(HH). This latter approach complements more direct determinations of theta from (3)J(HCOH) and potentially extends these more conventional analyses to O-substituted systems lacking the hydroxyl proton. (1)J(CH) values within hydroxymethyl fragments were also examined and found to depend on r(CH), which is modulated by specific bond orientation and stereoelectronic factors. These latter factors could be largely, but not completely, accounted for by C-C and C-O torsional variables, leading to only semiquantitative treatments of these couplings (details discussed in the Supporting Information). New equations pertaining to (2)J(HH) and (3)J(HH) have been applied to the analysis of hydroxymethyl group J-couplings in several mono- and oligosaccharides, yielding information on C5-C6 and/or C6-O6 rotamer populations.  相似文献   

8.
The occurrence of head-to-head (HH) addition in the radical polymerization of diallyl oxalate (DAO) was examined under various polymerization conditions. The content of HH linkage in poly (DAO) was reduced in comparison with allyl acetate and diallyl succinate; this may be ascribed to the high polarity of DAO inducing a polar effect on the intermolecular propagation of the growing polymer radical, resulting in reduced HH addition. The correlation between addition modes and evolution of carbon dioxide characteristic of DAO polymerization at elevated temperatures is mechanistically discussed in detail, with special focus on the solvent effect and the reduced dismutation of the cyclized radical compared to the uncyclized one.  相似文献   

9.
The heteropolynuclear complexes [Pd(2)M'(2)(mu-pz)(6)] (M'=Ag (1), Au (2); pzH=pyrazole), HT-[Pd(2)M'(2)(mu-3-tBupz)(6)] (M'=Ag (3 a), Au (4 a); 3-tBupzH=3-tert-butylpyrazole), and HH-[Pd(2)Au(2)(mu-3-tBupz)(6)] (4 b) have been prepared and some of them were structurally characterized. When 3-tert-butylpyrazolate was employed as a bridging ligand, two linkage isomers (head-to-tail (HT) and head-to-head (HH)) arise from the difference in orientation of the substituent groups on the pyrazolate bridges between the two Pd atoms. (1)H NMR spectroscopy has been used to identify and to follow the reversible stereochemical rearrangement of the HH isomer of [Pd(2)Ag(2)(mu-3-tBupz)(6)] (3 b) to form the HT isomer 3 a in CDCl(3) and the HT isomer of [Pd(2)Au(2)(mu-3-tBupz)(6)] (4 a) to form the HH isomer 4 b in C(6)D(6). Kinetic studies of the reaction have established the rate law to be -d(HH)/dt=d(HT)/dt=k(2)[HH]-k(1)[HT] for 3 b and -d(HT)/dt=d(HH)/dt=k(1)[HT]-k(2)[HH] for 4 a, where k(1) and k(2) denote the rate of isomerization from the HT to the HH isomer and that from the HH to the HT isomer, respectively. For typical runs at 50 degrees C in C(6)D(6), k(1)=13.8x10(-5) s(-1), k(2)=18.6x10(-5) s(-1), and K(eq)=k(2)/k(1)=1.24 for 3 b, and k(1)=1.26x10(-5) s(-1), k(2)=3.52x10(-5) s(-1), and K(eq)=k(1)/k(2)=0.36 for 4 a. Temperature-dependent rate measurements reveal DeltaH(not equal) and DeltaS(not equal) to be 100(1) kJ mol(-1) and 0(3) J mol(-1) K(-1) for 3 b and 112(5) kJ mol(-1) and 20(17) J mol(-1) K(-1) for 4 a, respectively. The rate of isomerization is essentially unaffected by the concentration of the complex or by the presence of neutral bridging ligands. These data and observations imply that the isomerization involves an intramolecular exchange process.  相似文献   

10.
11.
The comonomer composition and its distribution have been investigated for poly(3‐hydroxybutyrate‐co‐3‐hydroxyhexanoate) [P(3HB‐co‐3HH)], which was bacterially synthesized by Ralstonia eutropha from coconut oil as a carbon source. Using a chloroform/heptane mixed solvent, they were fractionated into several fractions with different comonomer composition. Bacterially synthesized P(3HB‐co‐3HH)s were found to have a wide compositional distribution. Using the fractions with a narrower comonomer composition distribution, the compositional dependence of thermal properties was investigated. The differential scanning calorimetry (DSC) melting behavior of a sample of unfractionated P(3HB‐co‐3HH) did not reflect that of fractions with similar average 3HH content. It was concluded that each of the fractions affects the thermal properties of the original unfractionated P(3HB‐co‐3HH), which should therefore be considered as polymer blends.  相似文献   

12.
The two-step synthesis of a new unsymmetrical ligand 2-[Ph2PC6H4C(H)=N]C6H4[N(H)COCH2N(H)CO2Bz], 2.HH, via acid-catalyzed Schiff base condensation of 2-(H2N)C6H4[N(H)COCH2N(H)CO2Bz], 1, with 2-Ph2PC6H4(CHO) in refluxing EtOH is reported. The multidentate ligand 2.HH, isolated in ca. 60% yield, exhibits an array of ligation modes, as exemplified by coordination studies with NiII, PdII, PtII, and AuI mononuclear metal precursors. Hence, reaction of 2.HH with AuCl(tht) (1:1 molar ratio, tht = tetrahydrothiophene) affords AuCl(2.HH), 3, in which the ligand behaves as a classic, neutral two-electron phosphorus donor. In contrast, reaction with MCl2(cod) (M = Pt, Pd; cod = cycloocta-1,5-diene) affords the corresponding dichloro complexes MCl2(2.HH) (4a M = Pt; 4b M = Pd) in which kappa2-P/N-chelation through both P and imino N-donor atoms is observed. Likewise, treatment of Pd(CH3)Cl(cod) with 2.HH gave Pd(CH3)Cl(2.HH), 4c, in which the imino nitrogen is trans to the methyl ligand. Cycloocta-1,5-diene elimination from, and single methyl protonation of, Pt(CH3)2(cod) with 1 equiv of 2.HH in toluene at ambient temperature affords the neutral complex Pt(CH3)(2.H-), 5a, in which 2.H- functions effectively in a kappa3-PNN' coordination mode. The dichloro compounds 4a or 4b undergo smooth N(H) deprotonation with tBuOK to give 6a\6a' and 6b\6b' in which 22- acts as a dianionic kappa4-PNN'N' ' ligand. The corresponding square-planar, diamagnetic, nickel(II) complex 6c\6c' was prepared in excellent yield from NiCl2.6H2O, 2.HH, and tBuOK. Variable-temperature NMR experiments confirm 6a\6a' and 6b\6b' exist, in solution, as a pair of conformational (anti and syn) isomers due to restricted rotation about the N-CO2Bz group. This feature is also borne out by single-crystal X-ray studies of anti-6a.CHCl3, syn-6a'.H2O, anti-6b.CHCl3, and anti-6c.CH2Cl2. To the best of our knowledge, we believe these constitute the first examples of crystallographically characterized conformers of a tetradentate ligand incorporating a P-donor center. All new compounds reported have been fully characterized by a combination of spectroscopic (NMR, FT-IR, ES-MS) and analytical methods. Furthermore, single-crystal X-ray studies have also been undertaken on compounds 2.HH, 3, 4a, and 5a.Et2O.  相似文献   

13.
Head-to-head bis(alpha-pyridonato)-bridged bis(ethylenediamine)dipalladium(ii), HH-[Pd(2)(en)(2)(alpha-pyridonato)(2)](ClO(4))(2), was synthesized and structurally characterized by X-ray crystallography. The (1)H NMR spectra show that the head-to-head (HH) dimer produces the head-to-tail (HT) dimer and monomers ([Pd(en)(alpha-pyridone)(2)](2+), [Pd(en)(H(2)O)(alpha-pyridone)](2+), [Pd(en)(H(2)O)(2)](2+), etc.) in aqueous solution, and the relative amount of dimers to monomers is dependent on the total concentration of the HH dimer dissolved as well as the acidity of the solution. It was found that the formation of the HH and HT dimers from the monomers is fast, and the HT dimer is produced from the HH dimer only via coexisting monomers, i.e., there is no direct isomerization path between the HH and HT dimers. The kinetic analyses for the HH <==>HT isomerization reaction with time-resolved (1)H NMR measurements revealed that the reaction proceeds via first-order kinetics, which was explained based on a relaxation process. The rate determining step for HH <==>HT isomerization is the reaction step between the mono-alpha-pyridone complex and the bis-alpha-pyridone complex, [Pd(en)(H(2)O)(alpha-pyridone)](2+)+alpha-pyridone <==> [Pd(en)(alpha-pyridone)(2)](2+).  相似文献   

14.
Early studies on cis-PtA2(d(G*pG*)) (A2 = diamine or two amines, G = N7-platinated G) and cis-Pt(NH3)2(d(G*pG*)) models for the key cisplatin-DNA cross-link suggested that they exist exclusively or mainly as the HH1 conformer (HH1 = head-to-head G bases, with 1 denoting the normal direction of backbone propagation). These dynamic models are difficult to characterize. Employing carrier A2 ligands designed to slow dynamic interchange of conformers, we found two new conformers, DeltaHT (head-to-tail G* bases with a Delta chirality) and HH2 (with 2 denoting the backbone propagation direction opposite to normal). However, establishing that the non-HH1 conformations exist as an intrinsic feature of the 17-membered Pt(d(G*pG*)) ring requires exploring a range of different carrier ligands. Here we employ the planar aromatic sp(2) N-donor 5,5'-Me(2)bipy (5,5'-dimethyl-2,2'-bipyridine) ligand, having a shape very different from those of previously used nonplanar sp(3) N-donor bidentate carrier ligands, which often bear NH groups. The 5,5'-Me(2)bipy H6 and H6' protons project toward the d(G*pG*) moiety and hinder the dynamic motion of 5,5'-Me(2)bipyPt(d(G*pG*)). We again found HH1, HH2, and DeltaHT conformers with typical properties, supporting the conclusions that the new DeltaHT and HH2 conformers exist universally in dynamic cis-PtA2(d(G*pG*)) adducts, including cis-Pt(NH3)2(d(G*pG*)), and that the carrier ligand typically has little influence on the overall structure of the Pt(d(G*pG*)) macrocyclic ring of a given conformer. The sizes of the G H8 to H6/H6' NOE cross-peaks indicate little base canting in all 5,5'-Me(2)bipyPt(d(G*pG*)) conformers, suggesting that carrier-ligand NH groups favor the canting of one G base in the HH1 and HH2 conformers of typical cis-PtA2(d(G*pG*)) adducts.  相似文献   

15.
We employ retro models, cis-PtA2G2 (A2 = a diamine, G = guanine derivative), to assess the cross-linked head-to-head (HH) form of the cisplatin-DNA d(GpG) adduct widely postulated to be responsible for the anticancer activity. Retro models are designed to have minimal dynamic motion to overcome problems recognized in models derived from cisplatin [A2 = (NH3)2]; the latter models are difficult to understand due to rapid rotation of G bases about the Pt-N7 bond in solution and the dominance of the head-to-tail (HT) form in the solid. Observation of an HH form is unusual for cis-PtA2G2 models. Recently, we found the first HH forms for a cis-PtA2G2 model with A2 lacking NH groups in a study of new Me2ppzPtG2 models. (Me2ppz, N,N'-dimethylpiperazine, has inplane bulk which reduces dynamic motion by clashing with the G O6 as the base rotates into the coordination plane from the ground state position approximately perpendicular to this plane G = 5'-GMP and 3'-GMP.) The finding of an HH form (albeit in a mixture with HT forms) with both G H8 signals unusually downfield encouraged us to study additional Me2ppzPtG2 analogues in order to explain the unusual spectral features and to identify factors that influence the relative stability of HT and HH forms. Molecular modeling techniques suggest HH structures with the H8's close to the deshielding region of the z axis of the magnetically anisotropic Pt atom, explaining the atypical shift pattern. When G = 1-Me-5'-GMP, we obtained NMR evidence that the HH rotamer has a high abundance (34%) and that the three rotamers have nearly equal abundance. These findings and the observation that the relative HT distributions varied little or not at all as a function of pH when G = Guo, 1-MeGuo, or 1-Me-5'-GMP are consistent with two of our earlier proposals concerning phosphate groups in HT forms of cis-PtA2(GMP)2 complexes. We proposed that a G phosphate group can form hydrogen bonds with the cis G N1H ("second-sphere" communication) and (for 5'-phosphate) A2 NH groups. The new results with 1-Me-5'-GMP led us to propose a new role for a 5'-phosphate group; it can also favor the HH form by counteracting the natural preference for the G bases to adopt an HT orientation. Finally, the HH form was also sufficiently abundant to allow observation of a distinct 195Pt NMR signal (downfield of the resonance observed for the HT forms) for several complexes. This is the first report of an HH 195Pt NMR signal for cis-PtA2G2 complexes.  相似文献   

16.
Depolarized light scattering is widely used to probe the spatial correlation of optical anisotropy in crystals, liquid crystals, and viscoelastic materials under stress, and a powerful means to study a non-equilibrium pattern evolution process of such a system. To follow the temporal change in the diagonal and off-diagonal contributions of the dielectric tensor, it is highly desirable to measure two-dimensional (2D) polarized (HH: horizontally transmitted, horizontally received) and depolarized (VH: vertically transmitted, horizontally received) scattering patterns simultaneously in a time-resolved manner. We develop a light scattering system with a video-rate time resolution as well as very high sensitivity to optical anisotropy. To detect extremely weak VH scattering from a sample without suffering from residual birefringence of the optical system itself and leakage of strong HH scattering signals, we use an objective lens specially designed for polarizing microscopy and Glan-laser prisms, respectively. This system enables us to experimentally elucidate the origin of VH scattering: we use the ratio of the VH and HH scattering intensity as a fingerprint for whether a 2D VH scattering pattern is caused by (i) optical anisotropy (intrinsic birefringence) or merely by (ii) spatial inhomogeneity of optically isotropic materials. We verify the validity of this method for a process of phase separation in a binary mixture of isotropic liquids. The simultaneous HH and VH measurement allows us to directly estimate the ratio of VH and HH scattering intensity accurately. The careful comparison of this ratio with a simple theory unambiguously demonstrates that the 2D VH scattering pattern is caused by the scattering angle dependence of the diffraction efficiency of light with the two polarization directions. That is, the origin of VH scattering is due to geometrical effects of the inhomogeneous distribution of the refractive index and not due to optical birefringence, as it should be for the optically isotropic sample. This method using the ratio of VH and HH scattering intensity may be widely used for distinguishing the two types of origins for a VH scattering pattern in an unambiguous manner.  相似文献   

17.
The multilayer multiconfiguration time-dependent Hartree (ML-MCTDH) method is discussed and a fully general implementation for any number of layers based on the recursive ML-MCTDH algorithm given by Manthe [J. Chem. Phys. 128, 164116 (2008)] is presented. The method is applied first to a generalized Henon-Heiles (HH) hamiltonian. For 6D HH the overhead of ML-MCTDH makes the method slower than MCTDH, but for 18D HH ML-MCTDH starts to be competitive. We report as well 1458D simulations of the HH hamiltonian using a seven-layer scheme. The photoabsorption spectrum of pyrazine computed with the 24D hamiltonian of Raab et al. [J. Chem. Phys. 110, 936 (1999)] provides a realistic molecular test case for the method. Quick and small ML-MCTDH calculations needing a fraction of the time and resources of reference MCTDH calculations provide already spectra with all the correct features. Accepting slightly larger deviations, the calculation can be accelerated to take only 7 min. When pushing the method toward convergence, results of similar quality than the best available MCTDH benchmark, which is based on a wavepacket with 4.6×10(7)time-dependent coefficients, are obtained with a much more compact wavefunction consisting of only 4.5×10(5) coefficients and requiring a shorter computation time.  相似文献   

18.
田真宁  许旋 《物理化学学报》2008,24(8):1482-1486
对PPh2py配合物[M(CO)3(PPh2py)2](M=Fe, Ru)的三种构型的异构体1-6进行了研究. 其中PPh2py以两个P原子与M配位形成HH构型1(Fe)和4(Ru), 以一个P和一个N原子与M配位形成HT构型2(Fe)和5(Ru), 以两个N原子与M配位形成HH’构型3(Fe)和6(Ru). 结果表明, (1) PPh2py中P原子对HOMO轨道的贡献最大, PPh2py作为电子给体时易以P原子与金属原子结合. (2)从分子能量和相互作用能数据表明, 配合物中HH构型最稳定, HH'构型最不稳定, 这与合成产物为HH构型的结果一致. (3) 键长和Wiberg键级均表明P—M键比N—M键结合力强. P、M原子间存在σ键, 而N、Fe原子间仅存在nN→n*M或nN→σ*M-P的电荷转移作用. (4) HH构型中M对HOMO的贡献最大, PPh2py向M的电荷转移最强, 使M的负电荷最大, 故HH构型最易作为电子给体以M原子与第二个金属配位形成双核配合物.  相似文献   

19.
The enthalpy of hydrophobic hydration, DeltaH(HH), was separately estimated for each apolar segment of amphiphiles and of related compounds. The specimens analyzed in this study were the homologs and analogs of sodium alkyl sulfate, alkyltrimethylammonium bromide, ammonium perfluoroalkylcarboxylate, and polyethylene glycols. The DeltaH(HH) was calculated from the adiabatic compressibility beta of aqueous solution systematically measured for each homologs at varying concentrations and temperatures. Those data reported in the previous paper were utilized. The DeltaH(HH) estimated for each -CH(2)-, -CF(2)-, and -OCH(2)CH(2)- segments in both hydrophobic and hydrophilic parts of the amphiphiles was found to change depending on its locations in the molecule. The DeltaH(HH) for the first -CH(2)- unit which is adjacent to the head sulfate group of H(CH(2))(n)SO(4)Na was found to be negligibly small; however, it increases gradually as the position of the unit moves apart from the anionic head group and attains as large as -26.5 kcal/mol for the eighth -CH(2)-. The change of DeltaH(HH) of -CF(2)- unit in F(CF(2))(n)COONH(4) shows the similar tendency to H(CH(2))(n)SO(4)Na. The value of DeltaH(HH) is as small as -8 kcal/mol for the first -CF(2)-, however, it increases to -16 kcal/mol for the fifth -CF(2)-, which is 1.5 times larger than that of the fifth -CH(2)- of alkyl sulfate. Contrary to this, all the DeltaH(HH) of the methylenes of cationic surfactant ranging from the second to the eighth -CH(2)- units of H(CH(2))(n)N(CH(3))(3)Br take a similar value around -8 kcal/mol of -CH(2)- irrespective of their positions. The first -CH(2)- unit which is equivalent to that of quaternary ammonium [HCH(2)](4)NBr shows slightly higher value of -9.0 kcal/mol of the first -CH(2)- than those for the second to eighth -CH(2)-. This seems to suggest that the cationic core nitrogen has an enhancing effect on hydrophobic hydration. The hydration enthalpy DeltaH(OEH) of each oxyethylene (OE) unit of polyethylene glycol was also estimated for H(OCH(2)CH(2))(p)OH (p=1-4) homologues to see the roles of ether oxygen and hydroxyl groups in hydrophobic hydration. The value of DeltaH(OEH) showed its maximum -13.1 kcal/mol for the OE unit located at the middle of three. The other two units at both ends were found to take the value of -9.9 kcal/mol, smaller than that at the center. This could suggest that the hydrophobic hydration of the ethylene segment is largely enhanced by the neighboring ether oxygens.  相似文献   

20.
Density functional theory (DFT) quantum chemical calculations of the structure and NMR parameters for highly strained hydrocarbon [2.2]paracyclophane 1 and its three derivatives are presented. The calculated NMR parameters are compared with the experimental ones. By least-squares fitting of the (1)H spectra, almost all J(HH) coupling constants could be obtained with high accuracy. Theoretical vicinal J(HH) couplings in the aliphatic bridges, calculated using different basis sets (6-311G(d,p), and Huz-IV) reproduce the experimental values with essentially the same root-mean-square (rms) error of about 1.3 Hz, regardless of the basis set used. These discrepancies could be in part due to a considerable impact of rovibrational effects on the observed J(HH) couplings, since the latter show a measurable dependence on temperature. Because of the lasting literature controversies concerning the symmetry of parent compound 1, D(2h) versus D(2), a critical analysis of the relevant literature data is carried out. The symmetry issue is prone to confusion because, according to some literature claims, the two hypothetical enantiomeric D(2) structures of 1 could be separated by a very low energy barrier that would explain the occurrence of rovibrational effects on the observed vicinal J(HH) couplings. However, the D(2h) symmetry of 1 with a flat energy minimum could also account for these effects.  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