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 109 毫秒
1.
α-Trifluoromethyl substituted allyl cations 3 and 4 have been prepared by ionizing their corresponding alcohols with SbF5 in SO2CIF at low temperatures. The barriers to rotation around the C1–C2 bond of the both cations were determined to be about 9 kcal/mol. The unusually low barriers as compared with their methyl analogues are rationalized by the unsymmetrical nature of the cations. CF3-substituted cyclohexenyl cations 8 and 10 were also prepared and characterized by 13C NMR spectroscopy. Density functional theory (DFT) calculations were performed to investigate geometries and charge densities of tifluoromethyl substituted allyl cations. 13C NMR chemical shifts of the cations were also calculated by IGLO method and compared with the experimental results.  相似文献   

2.
The homonuclear water-soluble and air stable compounds (dmpH) (H5O2) au][M(pydc)2].0.5H2O (M = Ni(II) (1), Cu(II) (2), Zn(II) (3); pydcH2 = pyridine-2,6-dicarboxylic acid, dipicolinic acid, dmp = 2,9-dimethyl-1,10-phenanthroline) have been prepared by self-assembly synthesis in aqueous solution at room temperature, and characterized by IR, 1H NMR, 13C NMR, elemental analysis and X-ray diffraction single crystal analyses for 1, 2 and 3. The complexes 1–3 represent the isostructural features. Extensive hydrogen bonding interactions involving all aqua ligands, dipicolinate oxygens and lattice water molecules further stabilize the complex units by linking them to form three dimensional polymeric networks. The stoichiometry and stability of the all three complexes in aqueous solution were investigated by potentiometric pH titration.  相似文献   

3.
A new azacrown bis-macrocycle (5) and its mono–cyclic analogue (7) were synthesized and characterized by FT-IR, 1H NMR, 13C NMR, DEPT 13C NMR, MS, and elemental analysis. The reaction with copper(II) nitrate yielded the corresponding complexes, formulated as Cu(5)(NO3)2·3H2O (8), and Cu(7)(NO3)2·2.5H2O (9). Also the stoichiometries of the complexes were determined in alcoholic solution and the results show that for both complexes the ratio of ligand to metal was 1:1 in methanol. The redox behavior of both complexes has been studied by cyclic voltammetry in DMF. Cyclic voltamogram of 8 shows quasi-reversible CuII/CuI redox couple whereas 9 shows a reversible CuII/CuI redox couple. Mono- and bis-macrocycle copper(II) complexes (8 and 9 respectively) cleaved plasmid pGS2 DNA by using an oxidative mechanism with 3- mercaptopropionic acid (MPA) as the reductant under aerobic conditions. The bis-macrocycle copper(II) complex 8 showed higher cleavage efficiency than their mono-macrocycle analogue 9 at the same Cu2+ concentration.  相似文献   

4.
Fast-atom bombardment (FAB) mass spectrometry was used to investigate the interaction of proton and alkali metal ions with dinucleotide analogs such as T-n-T (T = thymine moiety, n = polyether chain, e.g., triethylene, tetraethylene, pentaethylene, and hexaethylene ether 1–4), A-n-T (A = adenine unit 5–8), and T-n-OMe (9–12) in 3-nitrobenzyl alcohol matrix. The [M + H]+ ion is the most abundant ion for the A-n-T series, whereas in 1–4 and 9–12 the (TC2H4)+ ion is the most abundant. Formation of [M + H -C2H4O]+ ions, a characteristic fragmentation of crown ethers under electron ionization, is observed for compounds 1–12 and is more pronounced in 6 and 7. An abundant [M ? H]? ion is observed for all the compounds studied under negative ion FAB due to the presence of the (-CO-NH-CO-) group of thymine, an indication of existence of intramolecular H bonding. The FAB mass spectra of 1–12 with alkali metal ions (Li+, Na+, K+, Rb+, and Cs+) showed formation of abundant metal-coordinated ions ([M + Met]+ and [TC2H4 + Met]+). Compounds 3, 4, 6, 7, and 10–12 showed ions due to the substitution of the thymine moiety by a hydroxyl group ([M + Met ? 108]+, Met = metal ion). For compound 3 alone, substitution of two thymine groups ([M + Met - 216]+) was observed. Metastable ion studies were used to elucidate the structures of these potentially significant ions, and the ion formule were confirmed with high resolution measurements. Selectivity toward metal complexation with ligand size was seen in the T-n-T and A-n-T series and was even more pronounced in A-n-T series. These dinucleotide analogs fall in the following order of chelation of alkali metal ions, acyclic glymes < dinucleotide analogs (acyclic glymes substituted with nitrogen bases) < crown ethers, which places them in perspective as receptor models.  相似文献   

5.
The enantioselective interactions between chiral tetra-amidic receptors and nucleosides have been investigated by the ESI-IT-MS and ESI-FT-ICR-MS methodologies. Configurational effects on the CID fragmentation of diastereomeric [M H 2 ?H?A]?+ aggregates (A?=?2'-deoxycytidine dC, citarabine (ara-C) were found to be mostly offset by isotope effect in [S X 2 ?H?A]?+ (X?=?H, D) differently from the results obtained on the analogues (A?=?cytidine C and gemcitabine G). This result points the involvement of two different nucleoside/tetraamide isoforms. The structural differences of the [M H 2 ?H?A]?+ (A?=?C and G) complexes vs. the [M H 2 ?H?A]?+ (dC and ara-C) ones is fully confirmed by the kinetics of their uptake of the 2-aminobutane enantiomers, measured by FT-ICR mass spectrometry. Indeed, uptake of the 2-aminobutane enantiomers by [M H n ?H?A]?+ (n?=?1,2; A?=?dC and ara-C) complexes is reversible, while that by [M H n ?H?A]?+ (n?=?1,2; A?=?C and G) is not. The most encouraging result concerning the measured fragmentation and kinetic differences between C and ara-C, that are just epimers, indicates the possibility to subtly modulate the non-covalent drug/receptor interactions, through the electronic properties of the 2'-substituent on the nucleoside furanose ring, and furthermore on its three-dimensional position.  相似文献   

6.
Three novel coordination polymers, {[Co(L)(SO4)(H2O)](CH3OH)} (1), {[Cd(L)2(SiF6)](H2O)} (2) and [Zn(L)(NO3)2] (3), synthesized from 1,4-di(benzimidazole-1-yl)benzene (L), have been characterized by infrared spectroscopy, elemental analysis and single crystal X-ray diffraction. Compounds 1–3 exhibit different structures. Complex 1 has a 3-D diamond network containing 1-D CoII chains connected by SO 4 2? · anions; 2 has a 3-D-Po framework with 1-D porous channels along the c axis; and 3 has a 1-D zig-zag chain structure with a 2-D supramolecular network based on π-π interactions. The magnetic properties of 1 and the solid state fluorescence spectra of 2 and 3 have also been explored.  相似文献   

7.
A procedure for the synthesis of trans-Ru(NO)(Py)2Cl2(OH) (I) from K2[Ru(NO)Cl5] was proposed. Treatment of hydroxo complex I with HCl or H2SO4 at room temperature gave the corresponding salts trans-[Ru(NO)(Py)2Cl2(H2O)]Cl · 2H2O (II) and trans-[Ru(NO)(Py)2Cl2(H2O)]HSO4 (III). All the complexes obtained were characterized by 1H and 13C NMR and IR spectroscopy and elemental analysis; their structures were determined by X-ray diffraction. The structures are stabilized by π-stacking between the pyridine ligands of adjacent complex species.  相似文献   

8.
Multiple stage MS2 and MS3 mass spectrometric experiments, performed using a pentaquadrupole instrument, are employed to explore the gas-phase ion-molecule chemistry of several nitrilium [R-C≡N+-H (1), R-C≡N+-CH3 (2), and H-C≡N+-C2H5 (3)] as well as immonium ions RR1C=N+R2R3 (4) with the neutral diene isoprene. Polar [4+2+] Diels-Alder cycloaddition is observed for nitrilium ions when the energy gap between the lowest unoccupied molecular orbital (LUMO) of the ion and the highest occupied molecular orbital (HOMO) of the isoprene is small and the competing proton transfer reaction is endothermic. Thus, C-protonated methyl isonitrile H-C≡N+-CH3 (2a) and its higher homolog H-C≡N+-C2H5 (3a) form abundant [4+2+] cycloadducts with isoprene, but several protonated nitriles 1 do not; instead they show exothermic proton transfer as the main ion-molecule reaction. Replacement of the methyne hydrogen in 2a by a methyl, ethyl, or phenyl group (2b–d) raises the LUMO-HOMO gap, which greatly decreases the total yield of ion-molecule products and precludes cycloaddition. On the other hand, the electron-withdrawing acetyl and bromine substituents in 2e and 2f substantially lower the LUMO energy of the ions and cycloaddition reaction occurs readily. The simplest member of the immonium ion series, CH2=NH 2 + (4a), reacts readily by cycloaddition, whereas alkyl substitution on either the carbon or nitrogen (4b–f) dramatically lowers the overall reactivity, which substantially decreases or even precludes cycloaddition. In strong contrast, the N-phenyl (4g) and N-acetyl (4h) ions and the N-vinyl-substituted immonium ion, N-protonated 2-aza-butadiene (4i), react extensively with isoprene, mainly by [4+2+] cycloaddition. However, the isomeric C-vinyl-substituted ion (4j) displays only modest reactivity in both the proton-transfer and the cycloaddition channels. Collision-induced dissociation (CID) of the cycloadducts performed by on-line MS3 experiments demonstrates that they are covalently bound and supports their assignments as cycloaddition products. Retro Diels-Alder fragmentation is a major process for cycloadducts of both the immonium and the nitrilium ions, but other fragmentation processes also are observed. The cycloadduct of 4a with butadiene displays CID fragmentation identical to that of the authentic ion produced by protonation of 1,2,3,6-tetrahydropyridine, which thus strengthens the [4+2+] cycloaddition proposal. AM1 calculations also support the formation of the [4+2+] cycloadducts, which are shown in several cases to be much more stable than the products of simple addition, that is, the ring-open isomers.  相似文献   

9.
Photoionization mass spectrometry was used to investigate the dynamics of ion-neutral complex-mediated dissociations of the n-pentane ion (1). Reinterpretation of previous data demonstrates that a fraction of ions 1 isomerizes to the 2-methylbutane ion (2) through the complex CH3CH+CH 3 · CH2CH3 (3), but not through CH3CH+CH2CH 3 · CH3 (4). The appearance energy for C3Hin 7 + formation from 1 is 66 kJ mol?1 below that expected for the formation of n-C3H 7 + and just above that expected for formation of i-C3H 7 + . This demonstrates that the H shift that isomerizes C3H 7 + is synchronized with bond cleavage at the threshold for dissociation to that product. It is suggested that ions that contain n-alkyl chains generally dissociate directly to more stable rearranged carbenium ions. Ethane elimination from 3 is estimated to be about seven times more frequent than is C-C bond formation between the partners in that complex to form 2, which demonstrates a substantial preference in 3 for H abstraction over C-C bond formation. In 1 → CH3CH+CH2CH3 + CH3 by direct cleavage of the C1–C2 bond, the fragments part rapidly enough to prevent any reaction between them. However, 1 → 2 → 4 → C4H 8 + + CH4 occurs in this same energy range. Thus some of the potential energy made available by the isomerization of n-C4H9 in 1 is specifically channeled into the coordinate for dissociation. In contrast, analogous formation of 3 by 1 → 3 is predominantly followed by reaction between the electrostatically bound partners.  相似文献   

10.
Charge-transfer complexation of iodine with a new benzo-substituted macrocyclic diamide 5,6,7,8,9,10-hexahydro-2H-1,13,4,7,10-benzodioxatriazacyclopentadecine-3,11(4H,12H)-dione (L) with iodine was studied spectrophotometrically in chloroform, dichloromethane and their 1:1 (v/v) mixture. The observed time dependence of the charge-transfer band and subsequent formation of I3 - ion are related to the slow formation of the initially formed 1:1 L.I2 outer complex to an inner electron donor-acceptor (EDA) complex, followed by fast reaction of the inner complex with iodine to form a triiodide ion, as follows: L + I2L.I2 (outer complex), fast L.I2 (outer complex) → (L.I+)I- (inner complex), slow (L.I+)I- (inner complex) + I2 → (L.I+)I3 -, fast The pseudo-first-order rate constants for the transformation process were evaluated in different solvent systems. The stability constants of the resulting EDAr complexes were also evaluated and the solvent effect on their stability is discussed. The resulting complexes were isolated and characterized by FTIR and 1H NMR spectroscopy.  相似文献   

11.
Using cyclic voltammetry method, the reduction of cationic η6-fluorene complexes of manganese [(η6-9-R-C13H9)Mn(CO)2L]PF6 (L = CO, R = H (1 +); L = CO, R = CH3 (2 +); L = PnBu3, R = H (3 +); L = CO, R = tBu (4 +)) is studied. It is shown that, depending on the nature of a substituent in the position 9 of the fluorene ligand, the reduction occurs either with the detachment of an H atom from position 9 to give zwitterion compounds (complexes 1 ±, 2 ±, 3 ±) or with the attachment of an H atom into the coordinated ring of the fluorene ligand to given η5-cyclohexadienyl complex (η5-9-tBu-C13H9)Mn(CO)3 (5).  相似文献   

12.
By reacting calix[4]-1,3-aza-crown 2 with 1,4-phenyl diisothiocyanate in “1 + 2” condensation mode, the novel dumbbell shaped biscalix[4]-1,3-aza-crown 3 with rigid thiourea-bridge was conveniently prepared in yield of 85%. Its structure and conformation were characterized by elemental analyses, ESI-MS, 1H NMR and 1H–1H COSY techniques. The complexation properties of compound 3 were investigated by liquid–liquid extraction experiments and 1H NMR titration experiments. The results showed that compound 3 has excellent complexation abilities for soft cations and outstanding extraction selectivity for Ag+. The Ag+/Na+ and Ag+/Hg2+ extraction percentage of host 7 were as high as 43.2 and 16.9, respectively. 1H NMR titration experiments revealed the 1:2 stoichiometry of receptor 3-Ag+ complex was formed.  相似文献   

13.
Modification of [VO(OPri)3] with oximes in different molar ratios, yielded new class of vanadia precursors, [VO{OPri}3?n{L}n] {where, n = 1–3 and LH = C9H16C=NOH (1–3) and (CH3)2C=NOH (46)}.All the products are yellow in colour. (1) and (2) are liquid/viscous liquid, while others are solids. Molecular weight measurements of all these derivatives and the ESI-mass spectral studies of (1), (2), (3) and (5) indicate their monomeric nature. 1H and 13C{1H} NMR spectra suggest that the oximato moieties are monodentate in solution which was further confirmed by the 51V NMR signals, appeared in the region expected for tetra-coordinated oxo-vanadium atoms. On ageing, a disproportionation reaction occurs in (1) and some crystals appeared. Single crystal X-ray diffraction analyses of the crystals obtained from (1) as well as from (3) were found to be the same and indicate the presence of side-on {dihapto η 2-(N, O)} binding modes of the oximato ligands, leading to the formation of seven coordination environment around the vanadium atom. Thermogravimetric curve of (1) exhibits multi-step decomposition with the formation of V2O5 as the final product at ~850 °C. Sol–gel transformation of (3) yielded (a) VO2 sintered at 300 °C and (b) V2O5 at 600 °C. Similarly, sol–gel transformations of (1) and (2) yielded V2O5 (c) and (d) at 600 °C, respectively. Formation of monoclinic phase in (a) and orthorhombic phase in (b), (c) and (d) were confirmed by powder XRD patterns.  相似文献   

14.
Transition metal complexes of 2-(1-(carboxymethyl)-2-methyl-1H-benzimidazol-3-ium-3-yl)acetate (HL), namely [Co(L)2(H2O)4] · 6H2O (I) and [Cu(L)2(H2O)2] · 4H2O (II), have been synthesized by a hydrothermal procedure and characterized by X-ray crystallography, CIF files CCDC nos. 1007524 (I), 1007525 (II). Both I and II are mononuclear molecules. In I, the Co2+ ion is in octahedral coordiantion environment and surrounded by four O atoms from water molecules and two carboxylate O atoms of two deprotonated ligand (L?) occupied six culmination. While in II, the Cu2+ ion is located in a square-planar geometry, bounded to two aqua O atoms and two carboxylate O atoms from L?.  相似文献   

15.
o-Dihydroxy-3-(methylphenyl)-chromenones (coumarins; 3a3c) were synthesized from trimethoxybenzaldehydes through a reaction with the corresponding methylphenyl-acrylonitrile in pyridine·HCl and then H3O+. Dihydroxycoumarins reacted with the ditosylate or dichloride derivatives of tri- or tetraethyleneglycols in the presence of CH3CN/Na2CO3 and macrocyclic ethers with a coumarin moiety were obtained. The chromatographically purified new coumarin-crown ethers (5a–5f) were identified by IR, 1H NMR, and Mass spectrometry. The binding constants of Na+ with the coumarin-crown ethers were determined in an 80 % dioxane/water binary solvent system at 25 °C from conductance measurements.  相似文献   

16.
The energies of gas-phase solvation of the FSO3H 2 + , and ClSO3H 2 + cations and FSO 3 ? , ClSO 3 ? anions by one molecule of the respective acid (FSO3H or ClSO3H) have been calculated by a quantum chemical method taking into account electron correlation in terms of MP2 theory with a 6-311++G(d,p) basis set. The energy of additional solvation of the resulting complexes by liquid acid has been estimated within the continuum model of solvation by the IPCM method, with the acid modeled as a continuum with a large dielectric constant. The calculated data have provided a quantitative estimate for the energy of self-ionization of liquid FSO3H and ClSO3H acids. The similarly calculated energy of solvation of protons in 100% fluorosulfonic and chlorosulfonic acids is lower than the heat of hydration of the proton in aqueous solution by 23.4 and 24.5 kcal/mol, respectively. These quantum chemical data explain why 100% liquid fluorosulfonic and chlorosulfonic acids exhibit the properties of a superacid.  相似文献   

17.
The reaction of N-nitro-O-(4-nitrophenyl)hydroxylamine (1) with conc. H2SO4 affords 4-nitropyrocatechol and that with conc. sulfonic acids (RSO3H where R = Me, CF3) affords 2-hydroxy-5-nitrophenyl-R-sulfonates in yields of 80?C85%. These reactions are assumed to proceed through an intermediate (phenoxy)oxodiazonium ion [NO2C6H4O-N=N=O]+, which eliminates the N2O molecule to form the aryloxenium ion [NO2C6H4O]+. The latter reacts with acid anions at the ortho-carbon atom of the phenyl ring. The thermodynamical parameters of the elementary reactions resulting in the formation of the (phenoxy)oxodiazonium ion [NO2C6H4O-N=N=O]+ and aryloxenium ion [NO2C6H4O]+ were calculated in the B3LYP/6?311+G(d) study of the combined molecular system (nitrohydroxylamine 1 + [H3SO4]+). The reaction of nitrohydroxylamine 1 with aqueous solutions of strong acids (??70% H2SO4, CF3SO3H) affords mainly 4-nitrophenol. It appears that the mechanism of this reaction does not involve the formation of the aryloxenium ion.  相似文献   

18.
The chemical and electrochemical syntheses of the zinc (I) and cadmium (II) complexes are carried out on the basis of the tridentate Schiff base (H2L), the condensation product of (2-tosylaminoaniline) N-(2-aminophenyl)-4-methylbenzenesulfunamide with 1-phenyl-3-methyl-4-formylpyrazole-5-thiol. The structures and compositions of the synthesized metallochelates are proved by the data of C, H, and N elemental analyses, IR spectroscopy, and 1H NMR spectroscopy. X-ray absorption spectroscopy is used to determine the structure of the zinc complex. The binuclear structure of the cadmium complex is confirmed by the X-ray diffraction data (CIF file CCDC no. 1471159). The optical properties of H2L and the zinc and cadmium complexes in dimethyl sulfoxide (DMSO) solutions are studied.  相似文献   

19.
Photoionization was used to characterize the energy dependence of C3H 7 + , C3H 6 + , CH3OH 2 + and CH2=OH+ formation from (CH3)2)CHCH2OH+? (1) and CH3CH2CH2CH2OH+? (2). Decomposition patterns of labeled ions demonstrate that close to threshold these products are primarily formed through [CH 3 + CHCH3 ?CH2OH] (bd3) from 1 and through [CH3CH2CH2 ?CH2=OH+] (9) from 2. The onset energies for forming the above products from 1 are spread over 85 kJ mol?1, and are all near thermochemical threshold. The corresponding onsets from 2 are in a 19 kJ mol?1 range, and all except that of CH2=OH+ are well above their thermochemical thresholds. Each decomposition of 3 occurs over a broad energy range (> 214 kJ mol?1), This demonstrates that ion-permanent dipole complexes can be significant intermediates over a much wider energy range than ion-induced dipole complexes can be. H-exchange between partners in the complexes appears to be much faster than exchange by conventional interconversions of the alcohol molecular ions with their distonic isomers. The onsets for water elimination from 1 and 2 are below the onsets for the complex-mediated processes, demonstrating that the latter are not necessarily the lowest energy decompositions of a given ion when the neutral partner in the complex is polar.  相似文献   

20.
The reactions of 3(5)-(4-methoxyphenyl)-5(3)-phenyl-1H-pyrazole (L 1 ) with nitric acid and 5-(4-benzyloxyphenyl)-3-(furan-2-yl)-1H-pyrazole(L 2 ) with hydrochloric acid produced [HL 1 · NO3] (Salt-1) and [HL 2 · Cl] (Salt-2). The structures of Salt-1 and Salt-2 were determined by single crystal X-diffraction. In Salt-1, HL 1 showed [2 + 2] binding of NO3 ? ions in the solid state to form dimer architecture with R 1 2 (4) and R 4 4 (14) graph sets. An anion directed one-dimensional anion-assisted helical chain with active participation of the chloride ion and protonated pyrazole via N–H···Cl hydrogen bonding in Salt-2. In addition, the protonated HL 2 molecules interacted with each other through weak C–H···π interactions resulting in the formation of another one-dimensional helical chain.  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