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 31 毫秒
1.
The title compounds, C14H12N+·CH3O4S?, (I), and C15H14N+·CH3O4S?, (II), respectively, crystallize with the planar 10‐methylacridinium or 9,10‐di­methyl­acridinium cations arranged in layers, parallel to the twofold axis in (I) and perpendicular to the 21 axis in (II). Adjacent cations in both compounds are packed in a `head‐to‐tail' manner. The methyl sulfate anion only exhibits planar symmetry in (II). The cations and anions are linked through C—H?O interactions involving three O atoms of the anion, six acridine H atoms and the CH3 group on the N atom in (I), and the four O atoms of the anion, three acridine H atoms and the carbon‐bound CH3 group in (II). The methyl sulfate anions are oriented differently in the two compounds relative to the cations, being nearly perpendicular in (I) but parallel in (II). Electrostatic interaction between the ions and the network of C—H?O interactions leads to relatively compact crystal lattices in both structures.  相似文献   

2.
Hydration of alkylammonium ions under nonanalytical electrospray ionization conditions has been found to yield cluster ions with more than 20 water molecules associated with the central ion. These cluster ion species are taken to be an approximation of the conditions in liquid water. Many of the alkylammonium cation mass spectra exhibit water cluster numbers that appear to be particularly favorable, i.e., “magic number clusters” (MNC). We have found MNC in hydrates of mono- and tetra-alkyl ammonium ions, NH3(C m H2m+1)+(H2O) n , m=1–8 and N(C m H2m+1) 4 + (H2O) n , m=2–8. In contrast, NH2(CH3) 2 + (H2O) n , NH(CH3) 3 + (H2O) n1 and N(CH3) 4 + (H2O) n do not exhibit any MNC. We conjecture that the structures of these magic number clusters correspond to exohedral structures in which the ion is situated on the surface of the water cage in contrast to the widely accepted caged ion structures of H3O+(H2O) n and NH 4 + (H2O) n .  相似文献   

3.
Specific ion/molecule reactions are demonstrated that distinguish the structures of the following isomeric organosilylenium ions: Si(CH3) 3 + and SiH(CH3)(C2H5)+; Si(CH3)2(C2H5)+ and SiH(C2H5) 2 + ; and Si(CH3)2(i?C3H7)+, Si(CH3)2(n?C3H7)+, Si(CH3)(C2H5) 2 + , and Si(CH3)3(π?C2H4)+. Both methanol and isotopically labeled ethene yield structure-specific reactions with these ions. Methanol reacts with alkylsilylenium ions by competitive elimination of a corresponding alkane or dehydrogenation and yields a methoxysilylenium ion. Isotopically labeled ethene reacts specifically with alkylsilylenium ions containing a two-carbon or larger alkyl substituent by displacement of the corresponding olefin and yields an ethylsilylenium ion. Methanol reactions were found to be efficient for all systems, whereas isotopically labeled ethene reaction efficiencies were quite variable, with dialkylsilylenium ions reacting rapidly and trialkylsilylenium ions reacting much more slowly. Mechanisms for these reactions and differences in the kinetics are discussed.  相似文献   

4.
MINDO/3 calculations for singlet and triplet doubly charged benzene [C6H6]2+ are in satisfactory agreement with the experimentally determined values of the vertical double ionization energy of benzene; calculations for straight chain isomeric structures are consistent with the observed kinetic energy release on fragmentation to [C5H3]+ and [CH3]+. Symmetrical doubly charged benzene ions relax to a less symmetrical cyclic structure having sufficient internal energy to fragment by ring opening and hydrogen transfer towards the ends of the carbon chain. Fragmentation of [CH3C4CH3]2+ to [CH3C4]+ and [CH3]+ is a relatively high energy process (A), whereas both (B): [CH3CHC3CH2]2+ to [CHC3CH2]+ and [CH3]+ and (C): [CH3CHCCHCCH]2+ to [CHCCHCCH]+ and [CH3]+ may be exothermic processes from doubly charged benzene. Furthermore, the calculated energy for the reverse of process (A) is less than the experimentally observed kinetic energy released, whereas larger energies for the reverse of processes B and C are predicted. Heats of formation of homologous series [HCn]+, [CH3Cn]+, [CH2Cn?2CH]+, [CH3Cn?2CH2]+ and [CH2?CHCn?3CH2]+ with 1 < n < 6 are calculated to aid prediction of the most stable products of fragmentation of doubly charged cations. The homologous series [CH2Cn?2CH]+ is relatively stable and may account for ready fragmentation of doubly charged ions to [CnH3]+; alternatively the symmetrical [C5H3]+ ion [CHCCHCCH]+ may be formed. Dicoordinate carbon chains appear to be important stabilizing features for both cations and dications.  相似文献   

5.
《Tetrahedron》1986,42(22):6225-6234
Ab initio molecular orbital calculations on the distonic radical cations CH2(CH2)nN+H3 and their conventional isomers CH3(CH2)nNH2+ (n = 0,1, 2 and 3) indicate a preference in each case for the distonic isomer. The energy difference appears to converge with increasing n towards a limit which is close to the energy difference between the component systems CH3·H2+CH3+NH3 (representing the distonic isomer) and CH3CH3+CH3NH2+ (representing the conventional isomer). The generality of this result is assessed by using results for the component systems CH3·Y+CH3X+H and CH3YH+CH3X+. (or CH3YH+. + CH3X) to predict the relative energies of the distonic ions ·Y(CH2)nX+H and their conventional isomers HY(CH2)nX+. (X = NH2, OH, F, PH2, SH, Cl; Y = CH2, NH, O) and testing the predictions through explicit calculations for systems with n = 0,1 and 2. Although the predictions based on component systems are often close to the results of direct calculations, there are substantial discrepancies in a number of cases; the reasons for such discrepancies are discussed. Caution must be exercised in applying this and related predictive schemes. For the systems examined in the present study, the conventional radical cation is predicted in most cases to lie lower in energy than its distonic isomer. It is found that the more important factors contributing to a preference for distonic over conventional radical cations are the presence in the system of a group(X) with high proton affinity and the absence of a group (X, Y or perturbed (C—C) with low ionization energy.  相似文献   

6.
The reactions of ten metastable immonium ions of general structure R1R2C?NH+C4H9 (R1 = H, R2 = CH3, C2H5; R1 = R2 = CH3) are reported and discussed. Elimination of C4H8 is usually the dominant fragmentation pathway. This process gives rise to a Gaussian metastable peak; it is interpreted in terms of a mechanism involving ion-neutral complexes containing incipient butyl) cations. Metastable immonium ions ontaining an isobutyl group are unique in undergoing a minor amount of imine (R1R2C?NH) loss. This decomposition route, which also produces a Gaussian metastable peak, decreases in importance as the basicity of the imine increases. The correlation between imine loss and the presence of an isobutyl group is rationalized by the rearrangement of the appropriate ion-neutral complexes in which there are isobutyl cations to the isomeric complexes containing the thermodynamically more stable tert-butyl cations. A sizeable amount of a third reaction, expulsion of C3H6, is observed for metastable n-C4H9 +NH?CR1R2 ions; in contrast to C4H8 and R1R2C?NH loss, C3H6 elimination occurs with a large kinetic energy release (40–48 kJ mol?1) and is evidenced by a dish-topped metastable peak. This process is explained using a two-step mechanism involving a 1,5-hydride shift, followed by cleavage of the resultant secondary open-chain cations, CH3CH+ CH2CH2NHCHR1R2.  相似文献   

7.
The preparation of (borinato)(cyclobutadiene)cobalt complexes from the reactions of Co(C5H5BR)(1,5-C8H12) with acetylenes C2R′2 and of [C4(CH3)4]Co(CO)2I with Tl(C5H5BR) (R,R′ = CH3, C6H5) is described.In electrophilic substitution reactions Co(C5H5BCH3)[C4(CH3)4] (IVa) is more reactive than ferrocene. CF3CO2D effects H/D-exchange in the α-position of the borabenzene ring within a few minutes at ambient temperature and in the γ-position within less than four hours Friedel-Crafts acetylation with CH3COCl/AsCl3 in CH2Cl2 affords the 2-acetyl and the 2,6-diacetyl derivative of IVa. With the more active catalyst AlCl3, ring-member substitution is effected to give cations [Co(arene)C4(CH3)4]+ (arene = C6H5CH3, 2-CH3C6H4COCH3). Vilsmeier formylation gives the 2-formyl derivative of IVa. The acyl derivatives Co(2-R1CO-6-R2C5H3BCH3)[C4(CH3)4] (R1 = CH3, R2 = H, CH3CO and R1 = R2 = H) transform to the corresponding cations [Co(ortho-R1R2C6H4)C4(CH3)4]+ in superacidic media. The mechanistic relationship between acylation and ring-member substitution is discussed in detail.  相似文献   

8.
The gas-phase clustering reactions of proton in propanol and acetone, and chloride ions in acetone were investigated. The −ΔHn−1,n values obtained for clustering reactions (n−1,n) were as follows: H+ (C3H7OH)n−1 + C3H7OH ⇄ H+ (C3H7OH)n, (2, 3) 18.9 kcal mol−1, (3, 4) 14.2 kcal mol−1, (4, 5) 11.7 kcal mol−1; H+ (CH3COCH3)2 + CH3COCH3 ⇄ H+ (CH3COCH3)3, 14.2 kcal mol−1; and Cl + CH3COCH3 ⇄ Cl (CH3COCH3), 12.4 kcal mol.−1. For clustering reactions, Cl (CH3COCH3n−1 + CH3COCH3 ⇄ Cl (CH3COCH3)n where n ≥ 2, the equilibria could not be established; probably due to the isomerization of ligand acetone molecules from the keto to enol form.  相似文献   

9.
A series of novel cationic gemini surfactants, p-[C n H2n+1N+(CH3)2CH2CH(OH)CH2O]2C6H4·2Cl? [A(n = 12), B(n = 14) and C(n = 16)], containing a spacer group with two flexible and hydrophilic groups (2-hydroxy-1,3-propylene) on both sides of a rigid and hydrophobic group (1,4-dioxyphenylene) has been synthesized by the reaction of hydroquinone diglycidyl ether with N,N-dimethylalkylamine and N,N-dimethylalkylamine hydrochloride. Their surface-active properties have been investigated by surface tension measurement. The critical micelle concentration (cmc) values of the synthesized cationic gemini surfactants are one order of magnitude lower than those of their corresponding monomeric surfactants (C n H2n + 1N+(CH3)3·Cl?). Both the cmc and surface tension at the cmc (γcmc) of A are lower than those of p-[C12H25N+(CH3)2CH2]2C6H4·2Cl? (D). The novel cationic gemini surfactants A and B also show good foaming properties.  相似文献   

10.
(1,5-Cyclooctadiene) (4-substituted pyridinium 2-pyridylcarbonylmethylide)- rhodium(I) perchlorates, [Rh(COD)(C5H4NC(O)C?H+C5H4X-4)]ClO4 [COD = 1,5-cyclooctadiene; X = CH3C(O), CH3OC(O), C6H5, CH3, and H], have been prepared. They are shown to have the geometry with coordination by the pyridyl nitrogen and carbonyl oxygen atoms of the ylide ligands and to exhibit intramolecular rearrangement of coordinated COD in chloroform, methanol, and dimethyl sulphoxide based on IR and 1H NMR spectroscopies. Although the ylides have exhibited fluorescence bands due to an intramolecular charge-transfer transition and phosphorescence bands due to a carbonyl 3(n*) transition, the complexes have given emission bands due to the metal-to-ylide ligand charge-transfer transition. A.single crystal X-ray crystal structure has been determined for [Rh(COD)(C5H4NC(O)C?H+C5H4CH3-4)]ClO4. The crystals are monoclinic, space group P21/n with cell dimensions a = 14.887(3), b = 20.274(4), c = 6.966(1) Å, β = 96.13(1)°, and Z = 4. The structure has been refined by a block-diagonal least-squares method to final R = 0.060 for 2997 independent reflections with |Fo| > 3σ(F). The ylide carbon-pyridinium nitrogen bond distance is 1.420(10) Å. The bonded distances from rhodium to the midpoints of the double bonds of COD are 1.982(11) and 2.014(12) Å.  相似文献   

11.
Pyridine N-imine complexes of methylcobaloxime, CH3Co(Hdmg)2(R1— C5HnN+N?H) (n = 4; R1 = H, 2-CH3, 3-CH3, 4-CH3: n = 3; R1 = 2,6-CH3), have been synthesized by the reaction of CH3Co(Hdmg)2S(CH3)2 with a pyridine N-imine which is generated from a pyridine, hydroxylamine-O-sulfonic acid and K2CO3. The reactions of CH3Co(Hdmg)2(C5H5N+N?H) with acid anhydrides form new methylcobaloxime complexes with N-substituted pyridine N-imines, CH3Co(Hdmg)2(C5H5N+N?R2) R2 = COPh, COMe, COEt). With maleic anhydride, (pyridine N-acryloylimine)carboxylic acid is formed. With acetylenedicarboxylic acid dimethyl ester, 1,3-dipolar cycloaddition of the ligand gives pyrazolo[1,5-a]pyridine-2,3-dicarboxylic acid dimethyl ester.  相似文献   

12.
The effect of adding aliphatic alcohols (C4OH, C5OH, C6OH) and corresponding amines (C4NH2, C5NH2, C6NH2) on a series of dicationic gemini surfactants with the general formula C14H29(CH3)2N+?C(CH2)s?CN+(CH3)2C14H29, 2Br? (14-s-14; s=4,5,6), in the absence and presence of KNO3, has been studied by viscosity measurements at 303.15?K. As the chain length of the additive increased, the viscosity increased with increasing additive concentration and the extent of the effect followed the sequence: C6OH>C5OH>C4OH; C6NH2>C5NH2>C4NH2. The simultaneous presence of salt and additives showed an increase in ?? r values due to a synergistic effect. However, for equal chain lengths in the additives, the effect was greater for the n-alcohols. The tendency for the micelles to grow from spherical to rod-like structures is mainly influenced by the spacer chain length. At 303.15?K, the micellar growth was more pronounced for the shorter spacer, i.e. s being 4, which can be interpreted in terms of the short spacer having a higher tendency for micellar growth. Contrary to the cationic geminis, no effect was observed with a conventional surfactant of equal chain length, TTAB, even in the presence of KNO3 at the same concentration used for the geminis.  相似文献   

13.
The chiral cations, [CpFe(CO)(EMe2)L]+, are obtained both by reaction of [CpFe(CO)(EMe2)2]+ with the ligands (L) by heating, and by irradiation of the cations [C5H5Fe(CO)2EMe2]+ in the presence of L (E = S, Se, Te; L = PR3, AsR3, SbR3). The inversion about the chalcogen atom is investigated by DNMR spectrocopy. Compounds of the type [C5H5Fe(TeMe2)L2]+] are formed by irradiation of [C5H5Fe(CO)2(TeMe2)]+ and the ligands (L2 = 2 PR3, R = CH3, OCH3, OC6H5; L2 = R2P(CH2)nPR2, R = C6H5, n = 1,2,3). 77Se and 125Te NMR data vary according to the donor properties of the ligand L in the complexes.  相似文献   

14.
The effect of the addition of 2-methoxyethanol on the critical micelle concentration (cmc) and on the degree of counterion dissociation (??) of butanediyl-1,4-bis(tetradecyldimethylammonium bromide) gemini surfactant, [C14H29N+(CH3)2?C(CH2)4?CN+(CH3)2C14H29,2Br?] (referred as 14?C4?C14,2Br?), has been studied by varying the compositions of the 2-methoxyethanol + water mixed solvent media (0 to 50?%). To determine various thermodynamic parameters of micellization, on the basis of the mass?Caction model for micelle formation, the experiments were performed at selected compositions of the mixed solvent at four temperatures ranging between 25?°C and 50?°C. Furthermore, the air/bulk surface tensions of the pure and mixed media were determined, and a successful attempt was made to correlate the cohesive energy density described through the Gordon parameter with the values of Gibbs energy of micellization.  相似文献   

15.
N,N,N??,N??-tetramethylethylenediamine is obtained by the reaction of ethylenediamine with formaldehyde and formic acid (the Eschweiler-Clarke reaction) and then alkylated with allyl chloride (or bromide) in a ratio of 1:1 or 1:2 to obtain N-allyl-N,N,N??,N??-tetramethylethylenediaminium and N,N??-diallyl-N,N,N??,N??-tetramethylethylenediaminium bromide respectively. [{C2H4N2(H+)(CH3)4(C3H5)}Cu4Cl6] (1) and [{C2H4N2(CH3)4(C3H5)2}0.5Cu2Cl1.67Br1.33] (2) ??-complexes are obtained from alcohol solutions containing an ethylenediamine derivative and copper(II) chloride by ac-electrochemical synthesis on copper wire electrodes. An XRD study of the complexes is carried out. The crystals are monoclinic; 1: P21/n space group, a = 9.0081(6) ?, b = 12.5608(7) ?, c = 16.8610(10) ?, ?? = 102.061(3)°, V = 1865.7(2) ?3, Z = 4; 2: C2/c space group, a = 14.462(2) ?, b = 12.519(1) ?, c = 12.762(2) ?, ?? = 107.861(5)°, V = 2199.1(4) ?3, Z = 8. The structure of 1 consists of infinite copper halide networks with four crystallographically independent copper atoms, one of which coordinates the double bond of the allyl group of the ligand. The [C2H4N2(H)(CH3)4(C3H5)]2+ cations are attached above and below the plane of the network. The individual fragments are bonded via an extensive system of (N)H??Cl and (C)H??Cl hydrogen bonds. The structure of 2 contains a three-dimensional copper halide framework whose cavities contain the [C2H4N2(CH3)4(C3H5)2]2+ cations that are ??-coordinated with copper(I) atoms. In both structures, the Cu(I) atom that coordinates the C=C bond has a trigonal-pyramidal coordination environment consisting of the double C=C bond of the corresponding ligand and three halogen atoms. The other Cu(I) atoms have a tetrahedral environment consisting solely of halogen atoms. The Cu-(C=C) distance is 1.958(1) ?, (1) and 1.974(1) ? (2).  相似文献   

16.
Reactions of reactive cyclopentadienyliron complexes C5H5Fe(CO)2I, [C5H5Fe(CO)2THF]BF4, [C5H5Fe(CO)((CH3)2S)2]BF4 and [C5H5Fe(p-(CH3)2C6H4)]PF6 with P(OR)3 as ligands (R = CH3, C2H5, i-C3H7 and C6H5) lead to the formation of the complex compounds C5H5Fe(CO)2?n(P(OR)3)nI and [C5H5Fe(CO)3?n(P(OR)3)n]X (n = 1, 2 and n = 1–3, X = BF4, PF6). Spectroscopic investigations (IR, 1H, 13C and 31P NMR) indicate an increase of electron density on the central metal with increasing substitution of CO groups by P(OR)3 ligands. The stability of the compounds increase in the same way.  相似文献   

17.
Reactions that proceed within mixed ethylene–methanol cluster ions were studied using an electron impact time-of-flight mass spectrometer. The ion abundance ratio, [(C2H4)n(CH3OH)mH+]/[(C2H4)n(CH3OH)m+], shows a propensity to increase as the ethylene/methanol mixing ratio increases, indicating that the proton is preferentially bound to a methanol molecule in the heterocluster ions. The results from isotope-labelling experiments indicate that the effective formation of a protonated heterocluster is responsible for ethylene molecules in the clusters. The observed (C2H4)n(CH3OH)m+ and (C2H4)n(CH3OH)m–1CH3O+ ions are interpreted as a consequence of the ion–neutral complex and intracluster ion–molecule reaction, respectively. Experimental evidence for the stable configurations of heterocluster species is found from the distinct abundance distributions of these ions and also from the observation of fragment peaks in the mass spectra. Investigations on the relative cluster ion distribution under various conditions suggest that (C2H4)n(CH3OH)mH+ ions with n + m ≤ 3 have particularly stable structures. The result is understood on the basis of ion–molecule condensation reactions, leading to the formation of fragment ions, $ {\rm CH}_2=\!=\mathop {\rm O}\limits^ + {\rm CH}_3 $ and (CH3OH)H3O+, and the effective stabilization by a polar molecule. The reaction energies of proposed mechanisms are presented for (C2H4)n(CH3OH)mH+(n + m ≤ 3) using semi-empirical molecular orbital calculations.  相似文献   

18.
Bulky phosphanes PR3 (R = C6H11, iC3H7, t-C4H9, C6H4CH3-o) stabilize complexes of type [C5H5Ni(PR3)L]BF4 (L=S(CH3)2, (CH3)3PS), from which [C5H5Ni(PR3)2]+ cations are obtained. Iodide replaces the sulfur ligands to yield neutral C5H5Ni(PR3)I compounds. No stable [C5H5Ni(PR3)]+ cations could be obtained by iodide abstraction, but [C5H5Ni(PR3)CO]+ cations were formed in the presence of carbon monoxide.  相似文献   

19.
The thermodynamic quantities Kn?1 n, ΔG0n?1, n and ΔS0n?1, n for the gas phase equilibrium reactions RNH+3(RNH2)n?1 + RNH2 = RNH+3(RNH2)n, where n ? 3 and R indicates an alkyl group (CH3, C2H5, n-C3H7 and iso-C3H7), have been determined.  相似文献   

20.
Fifteen [C6H5(X)FeCp]+ cations with substituents X having different electron-donating or electron-withdrawing effects were treated with NaBH4 in glyme or THF. The relative distributions of products from o-, m-, p- and ipso-additions of the hydride ion to the arene ring were determined by high resolution 1H NMR. For the η6-N,N-dimethylaniline-η5-cyclopentadienyliron cation with the most electron-donating of the substituents studied, only m- and p-hydride addition products were obtained, while in the reaction with the η6-nitrobenzene-η5-cyclopentadienyliron cation, which contained the most electron-withdrawing of the substituents investigated, only the o-addition product was formed. For the other 13 cations, with X  C6H5O, CH3O, p-CH3C6H4S, C6H5CH2, (CH3)3C, CH3, CH3CH2, C6H5, Cl, COOCH3, C6H5CO, CN and p-CH3C6H4SO2, o-, m- and p-hydride addition products were obtained in all cases, with a few instances also giving very minor amounts of ipso-adducts. The relative product distributions observed were interpreted by suggesting that while electronic effects played a major role, steric factors and free valency effects favoring o-addition as suggested by MO calculations [5] could also exert their influence in giving rise to the overall results.  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