首页 | 官方网站   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.
Complexation of K+ by 18-crown-6 ether (18C6) in pure water and in acetonitrile–water mixed solvents containing 0.1 mol-dm? 3 (C2H5)4NCl has been systematically studied by isothermal titration calorimetry (ITC) at 293, 298, and 303 K. The formation constant K of the 1:1 [K(18C6)]+ complex and the complexation enthalpy Δ rH were simultaneously determined from the titration data. The logK and Δ rH(kJ-mol? 1) values at 298 K are 2.04, ?26.2 in pure water and 2.23, ?25.0; 2.61, ?24.2; 2.95, ?23.8; 3.48, ?21.0; 3.85, ?19.4; 4.36, ?18.7; and 5.73, ?17.0 in the mixed solvents at x AN (mole fraction of acetonitrile) of 0.043, 0.135, 0.258, 0.448, 0.578, 0.759, and 1.0, respectively. The change in heat capacity for the complex formation, Δ C p °, was also determined by the temperature dependence of Δ rH. Whereas the Δ C p ° is (57 ± 11) and (63 ± 20) J-mol? 1-K? 1 in pure water and in the solvent mixture at x AN = 0.043, respectively, it decreases with increasing x AN. The Δ C p ° values are ?(48 ± 11), ?(110 ± 25), ?(354 ± 40), ?(359 ± 24), and ?(304 ± 30) J-mol? 1-K? 1 at x AN = 0.135, 0.258, 0.448, 0.578, and 0.759, respectively. The changes in complexation thermodynamics (Δ Δ rG, Δ ΔrH, and Δ Δ r S) are discussed in terms of the corresponding transfer thermodynamics of K+, 18-crown-6, and [K(18C6)]+ upon transferring from water to acetonitrile–water mixed solvents. It was found that hydrophobic solvation of the complex [K(18C6)]+ plays an important role in complex formation occurring in water and in the water-rich mixed solvent. Moreover, changes in solvent structure significantly affect the transfer enthalpy and entropy of each species, i.e., K+, 18-crown-6, and [K(18C6)]+. The observed monotonous changes in the complexation Gibbs energy, enthalpy, and entropy with solvent composition are due to the effective compensation of the Δ trG, Δ trH, and Δ tr S for K+ with those for 18-crown-6 and [K(18C6)]+.  相似文献   

2.
Catalysis of the reaction between Me3SiO[Si(H)Me)O]nSiMe3 (n ≈ 50) and ethanol by silica-supported rhodium and iridium catalysts has been investigated. Donor groups in the anchored ligands were PPh2, S?, or C5H4?. The rhodium-PPh2 system showed marked inhibition by dihydrogen. The supported iridium catalysts all showed high activity which declined rapidly during successive cycles of re-use, but the iridium-PPh2 catalyst was the least affected. In every case, the separated liquid products showed activity as homogeneous catalysts, indicating that leaching of the metal from the support was occurring. That ligand was also being leached was shown by labelling with tritium. The results demonstrate the necessity to test supported catalysts through more than one cycle.  相似文献   

3.
Ab initio RHF calculations with the 3–21G basis set have been carried out on cycloadditions of CF2 and CCl2 with ethylene. Although π-complex intermediates are predicted for both reactions at this level, MP2/3-21G calculations imply that there are no complexes in reactions of CCl2 or more reactive carbenes with ethylene or substituted alkenes. Nevertheless, negative activation energies can be observed, since ΔG reaches a maximum due to the increase in —TΔS for these bimolecular reactions. The apparent “entropy control” for reactive carbenes results from the rapid decrease in ΔH. As the reactivity of the alkene increases, the transition state shifts to an earlier point on the free energy surface, where —TΔS3 is lower, but ΔH3 is higher than for less reactive alkenes. Model potentials are developed for ΔH and —TΔS which reproduce experimental behavior, without the assumption of intermediates.  相似文献   

4.
For decades, high-resolution 1H NMR spectroscopy has been routinely utilized to analyze both naturally occurring steroid hormones and synthetic steroids, which play important roles in regulating physiological functions in humans. Because the 1H signals are inevitably superimposed and entangled with various JH–H splitting patterns, such that the individual 1H chemical shift and associated JH–H coupling identities are hardly resolved. Given this, applications of thess information for elucidating steroidal molecular structures and steroid/ligand interactions at the atomic level were largely restricted. To overcome, we devoted to unraveling the entangled JH–H splitting patterns of two similar steroidal compounds having fully unsaturated protons, i.e., androstanolone and epiandrosterone (denoted as 1 and 2, respectively), in which only hydroxyl and ketone substituents attached to C3 and C17 were interchanged. Here we demonstrated that the JH–H values deduced from 1 and 2 are universal and applicable to other steroids, such as testosterone, 3β, 21-dihydroxygregna-5-en-20-one, prednisolone, and estradiol. On the other hand, the 1H chemical shifts may deviate substantially from sample to sample. In this communication, we propose a simple but novel scheme for resolving the complicate JH–H splitting patterns and 1H chemical shifts, aiming for steroidal structure determinations.  相似文献   

5.
The first synthesis of the naturally occurring cyclic peptide axinellin A has been achieved. Cyclization and subsequent deprotection of linear precursors containing either a t-butyl protected Thr residue or a Thr(ΨMe,Mepro) derivative gave a cyclic peptide, identical in all respects to the naturally occurring material, with the exception that the synthetic peptide does not exhibit the cytotoxic activity reported for the natural product.  相似文献   

6.
Because electrophiles regulate many signalling pathways in cells, by modifying cysteine residues in proteins, they have a wide range of biological activity. In this study, a deuterium-labelling mass spectrometry–tandem diode-array detector (MS–DAD) screening method was established for rapid discovery of naturally occurring electrophiles. Glutathione (GSH) was used as a probe and incubated with natural product extracts. To distinguish different types of electrophile, incubation was performed in two reaction solvents, H2O and D2O. Ten types of naturally occurring electrophile were chosen, on the basis of their properties, to undergo the screening assay. By using this screening method, we successfully discovered the bioactive electrophile 4-hydroxyderricin in an ethanol extract of Angelica keiskei. This electrophile had potent NAD(P)H:quinone oxidoreductase 1 (NQO1)-inducing activity at a concentration of 20 μmol L−1.  相似文献   

7.
Ab initio calculations at the STO—3G and 4—31G levels have been carried out for the H2SO4 molecule as a function of the pair of twist angles of the HO bonds about the respective OS bonds. Values for the remaining bond angles and lengths were taken from the recent microwave structural determination by Kuczkowski et al. The results indicate a minimum energy for a structure with a (sc, sc) conformation and C2 symmetry, where sc denotes synclinal, or gauche. This structure corresponds to that observed. At a higher energy of 11.5 kJ mol?1 (4—31G) there is a structure with a (+sc, ?sc) conformation and Cs symmetry. The torsional modes corresponding to the a and b irreducible representations of the C2 point group are estimated to have frequencies of 280 and 265 cm?1, respectively.  相似文献   

8.
Hydrogen-deficient peptide radical cations exhibit fascinating gas phase chemistry, which is governed by radical driven dissociation and, in many cases, by a combination of radical and charge driven fragmentation. Here we examine electron capture dissociation (ECD) of doubly, [M + H]2+?, and triply, [M + 2H]3+?, charged hydrogen-deficient species, aiming to investigate the effect of a hydrogen-deficient radical site on the ECD outcome and characterize the dissociation pathways of hydrogen-deficient species in ECD. ECD of [M + H]2+? and [M + 2H]3+? precursor ions resulted in efficient electron capture by the hydrogen-deficient species. However, the intensities of c- and z-type product ions were reduced, compared with those observed for the even electron species, indicating suppression of N?CC?? backbone bond cleavages. We postulate that radical recombination occurs after the initial electron capture event leading to a stable even electron intermediate, which does not trigger N?CC?? bond dissociations. Although the intensities of c- and z-type product ions were reduced, the number of backbone bond cleavages remained largely unaffected between the ECD spectra of the even electron and hydrogen-deficient species. We hypothesize that a small ion population exist as a biradical, which can trigger N?CC?? bond cleavages. Alternatively, radical recombination and N?CC?? bond cleavages can be in competition, with radical recombination being the dominant pathway and N?CC?? cleavages occurring to a lesser degree. Formation of b- and y-type ions observed for two of the hydrogen-deficient peptides examined is also discussed.  相似文献   

9.
2β,3α-Dihydroxyurs-12-en-28-oic acid(6) is a naturally occurring diastereoisomer of corosolic acid with glycogen phosphorylase inhibitory activity.A new strategy for the semi-synthesis of 6 was developed.Using the commercially available ursolic acid(1) as the starting materials,6 was synthesized through five facile reactions with a high stereoselectivity and an overall yield of 47.3%.The structure of 6 was confirmed by optical rotation.ESI-MS,1H NMR and 13C NMR data.  相似文献   

10.
Thorikosite, (Pb3Sb0.6As0.4)(O30H)Cl2, is a naturally occurring member of the bismuth oxyhalide group isostructural with LiBi3O4Cl2. The space group isI4/mmm witha = 3.919(1)A?,c = 12.854(5)A?, andZ = 1. A crystal structure analysis showed complete solid solution of Pb2+, Sb3+, and As3+ on the single cation site and large atomic temperature factors indicative of pervasive structural disorder. The latter is due to the structural adjustments necessary to accommodate cations of very different sizes in the same site. Thorikosite is closely related to synthetic tetragonal PbSbO2Cl through the coupled substitution Sb3+O2? ? Pb2+(OH)?.  相似文献   

11.
Stereoselective preparative enzymatic acylation and hydrolysis/methanolysis of various C-substituted rac-thiazol-2-yl-methanols were achieved for the preparation of enantiopure or enantiomerically enriched, naturally occurring 2-hydroxymethylthiazoles. The absolute configurations of the resulting secondary alcohols were determined by a detailed 1H NMR study of Mosher’s derivatives.  相似文献   

12.
Host?guest complexation has been studied by 1H NMR on the benzyl and phenethyl amides of ferulic and caffeic acids as the guests in chloroform and acetonitrile; the counter host is a cyclophane which integrates four phenylene rings, amino and amide groups in the macrocyclic framework and bears four pendant methyl acetate ester arms. CAPE, one of the best known natural antioxidants, also has been studied for comparison. Among the guests studied, ferulic acid benzyl amide shows NMR shifts due to the formation of a host?guest complex in chloroform. The complexation occurs in two steps with the formation constants K 1?=?[HG]/[H][G]?=?6?M?1 and β 2?=?[HG2]/[H][G]2?=?87?M?2. Two guest molecules are bound on the surface of the macrocyclic framework of a host molecule by two hydrogen bonds, NH(host amide)···O=C(guest amide) and C=O(host ester)···HO(guest phenol). The latter hydrogen bond may protect the bioactive site, i.e., phenol OH, of guest molecules captured in the complex against undesirable oxidation. This feature is observed only for ferulic acid benzyl amide in chloroform; the cyclophane ester interacts with this amide, distinctively from the other hydroxycinnamic acid derivatives.  相似文献   

13.
The nuclear spin—spin coupling constants J(C,H) and J(C,D) have been measured over the temperature range 200–370 K for the methane isotopomers 13CH4, 13CH3D, 13CHD3 and 13CD4. The coupling constants increase with increasing temperature for any one isotopomer and decrease with increasing secondary deuterium substitution at any one temperature. The results are entirely attributable to intramolecular effects and the data have been fitted by a weighted least-squares regression analysis to a spin—spin coupling surface thereby yielding a value for 1Je(C,H), the coupling constant at equilibrium geometry, and values for the bond length derivatives of the coupling. We find that 1Je(C,H) = 120.78 (±0.05) Hz which is about 4.5 Hz smaller than the observed value in 13CH4 gas at room temperature. Results are also reported for J(H,D) in 13CH3D and 13CHD3 for which no temperature dependence was detected.  相似文献   

14.
The reactions occurring between the components of metallocene and methylaluminoxane (MAO) catalyst leading to the reduction of the former were studied by electron paramagnetic resonance (EPR). At low Al/Zr ratios, CpZrCl3 (Cp = η5-cyclopentadienyl) was reduced to simple trivalent Zr species (g = 1.998, a(91Zr) = 12.3 G) without other superhyperfine splittings. At higher Al/Zr ratios the reactions proceed further to form two CpZr(III) hydrides (g = 1.991, a(H) = 5.5 G; and g = 2.00, a(H) = 3 G). Two CpTi(III) hydrides were also produced by the reaction of MAO with CpTiCl3 (g = 1.989, a(H) = 7.4 G, a(Ti) = 8 G; and, g = 1.995, a(H) = 4.5 G, a(Ti) = 8 G). In the case of Cp*TiCl3 (Cp* = η5-pentamethyl cyclopentadienyl) initially a multitude of paramagnetic species were formed. After long reaction time the final products show EPR features consistent with two η3: η4-(1,2,3-trimethyl-4,5-dimethylene cyclopentadienyl)hydrido Ti(III) species: the abundant one with g = 1.999, (H, sextet) = 9.5 G, a(Ti) = 9.5 G, and a weaker one of g = 1.975, a(H) = 4.8 G. All the five protons of these species and as well as those in the Cp hydrido complexes of Ti and Zr undergo facile H? D exchanges with D2. MAO is important in the formation of these hydrides because they are not formed by trimethyl aluminum reduction. The presence of tetrahydrofuran suppresses the hydride formation. The possible structures for the hydrido species and reactions producing them are discussed.  相似文献   

15.
In our previous paper, we reported that amphiphilic Ir complex–peptide hybrids (IPHs) containing basic peptides such as KK(K)GG (K: lysine, G: glycine) (e.g., ASb-2) exhibited potent anticancer activity against Jurkat cells, with the dead cells showing a strong green emission. Our initial mechanistic studies of this cell death suggest that IPHs would bind to the calcium (Ca2+)–calmodulin (CaM) complex and induce an overload of intracellular Ca2+, resulting in the induction of non-apoptotic programmed cell death. In this work, we conduct a detailed mechanistic study of cell death induced by ASb-2, a typical example of IPHs, and describe how ASb-2 induces paraptotic programmed cell death in a manner similar to that of celastrol, a naturally occurring triterpenoid that is known to function as a paraptosis inducer in cancer cells. It is suggested that ASb-2 (50 µM) induces ER stress and decreases the mitochondrial membrane potential (ΔΨm), thus triggering intracellular signaling pathways and resulting in cytoplasmic vacuolization in Jurkat cells (which is a typical phenomenon of paraptosis), while the change in ΔΨm values is negligibly induced by celastrol and curcumin. Other experimental data imply that both ASb-2 and celastrol induce paraptotic cell death in Jurkat cells, but this induction occurs via different signaling pathways.  相似文献   

16.
The FT-microwave spectrum of n-butylgermane, CH3CH2CH2CH2GeH3 has been investigated from 4000 to 18,000 MHz and the microwave spectra have been observed for all of the five naturally occurring germanium isotopologues for the anti-anti (aa) conformer. The dipole moment for the 74Ge containing species has been measured, giving a total dipole moment of 0.881 (26) D. In addition, the vibrational spectrum of n-butylgermane is described. Modestly complete assignments are made for the aa conformer. The relative stabilities of the five conformers are calculated, and the anti-anti (aa) conformer is found to be the most stable in all calculations done. This conclusion is confirmed by the infrared and Raman spectrum of the annealed crystal. The dipole moments of all conformers are calculated to be approximately equal and less than 1 D, ranging from approximately 0.8 to 0.9 D.  相似文献   

17.
Ab initio MO calculations using 6-31G and 6-31 + G (for complexes with F and LiF) basis sets have been carried out for complexes of H2O (monomer and dimer) with F, Cl, Li+ ions as well as with LiF and LiCl ion pairs for the evaluation of the OH stretching force constants. The changes in force constants are discussed in terms of molecular interactions, cooperativity effect and interionic electrostatic interactions. It is noticed that the cooperativity effect also operates through ionic bonds in symmetrically hydrated ion pairs and that OH stretching force constants are found to increase in the case of solvent bound ion pairs and symmetrically hydrated halide ions showing anticooperativity effect.  相似文献   

18.
19.
In this study, quantum chemical calculations of geometric parameters, conformational, natural bond orbital (NBO) and nonlinear optical (NLO) properties, vibrational frequencies, 1H and 13C NMR chemical shifts of the title molecule [C9H7F5N2O3] in the ground state have been calculated with the help of Density Functional Theory (DFT-B3LYP/6-311++G(d,p)) and Hartree-Fock (HF/6-311++G(d,p)) methods. The optimized geometric parameters, vibrational frequencies, 1H and 13C NMR chemical shifts values are compared with experimental values of the investigated molecules. Comparison between experimental and theoretical results showed that B3LYP/6-311++G(d,p) method is able to provide more satisfactory results. In order to understand this phenomenon in the context of molecular orbital picture, we examined the molecular frontier orbital energies (HOMO, HOMO-1, LUMO, and LUMO + 1), the energy difference (ΔE) between E HOMO and E LUMO, electronegativity (χ), hardness (η), softness (S) calculated by HF/6-311++G(d,p) and B3LYP/6-311++G(d,p) levels. The molecular surfaces, Mulliken, NBO, and Atomic polar tensor (APT) charges of the investigated molecule have also been calculated by using the same methods.  相似文献   

20.
Ti0.33V0.67HxDy (x+y≈0.9) alloys have been studied by means of X-ray powder diffraction (XRD), differential scanning calorimetry (DSC) and 1H and 2H NMR. The crystal structures are body-centered-cubic (bcc) dominantly, being mixed with a face-centered-cubic (fcc) phase. A phase transition similar to that from the δD phase to the αD phase in the V-D system is observed in all the samples except for the protide. H and D are considered to occupy tetrahedral sites. The temperature and frequency dependence of spin-lattice relaxation times T1 of 1H and 2H has been analyzed by Bloembergen-Purcell-Pound equations with a distribution of correlation times, and parameters of hydrogen diffusion are estimated. The mean activation energy for D diffusion (ED) is higher than that for H diffusion (EH). EH is constant while ED increases slightly with the [D]/[H] ratio. The distribution of the correlation times increases as the [D]/[H] ratio decreases.  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-23

京公网安备 11010802026262号