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 156 毫秒
1.
The crystals of a new melaminium salt, 2,4,6‐tri­amino‐1,3,5‐triazine‐1,3‐diium bis(4‐hydroxy­benzene­sulfonate) dihydrate, C3H8N62+·2C6H5O4S?·2H2O, are built up from doubly proton­ated melaminium(2+) residues, dissociated p‐phenol­sulfonate anions and water mol­ecules. The doubly protonated melaminium dication lies on a twofold axis. The hydroxyl group of the p‐hydroxybenzenesulfonate residue is roughly coplanar with the phenyl ring [dihedral angle 13 (2)°]. A combination of ionic and donor–acceptor hydrogen‐bond interactions link the melaminium and p‐hydroxybenzenesulfonate residues and the water mol­ecules to form a three‐dimensional network.  相似文献   

2.
Mass-analysed ion kinetic energy spectra for collisional activation (CA) of [C6H6]+˙ formed via electron capture by [C6H6]2+ ions in collision with neutral benzene molecules have been compared for the C6H6 isomers benzene, 1,5-hexadiyne and 2,4-hexadiyne. Comparisons of fragment abundance and total CA fragment yields were also made for [C6H6]+˙ ions generated by electron ionization (EI). CA conditions of ion velocity and collision gas pressure were identical in these comparisons. In general the fragment abundance patterns for the ions formed by charge exchange were very similar to those for singly charged benzene ions generated by EI. However, significant variations in CA fragment yield (the ratio of the total CA fragment ion abundance to the abundance of the incident unfragmented ions) were observed. It is not clear from the results whether these variations reflect structurally different ions or ions of different internal energies. The CA spectra of [C6H6]+˙ ions derived from charge exchange reactions between the benzene dication and the target gases He, Ne, Ar, Kr and Xe have also been recorded and, once again, very similar fragment abundance patterns were observed along with large variations in total CA fragment yields. Charge exchange efficiency measurements are reported for reactions between the benzene dication and the targets He, Ne, Ar, Kr, Xe and C6H6 (benzene) and also for the doubly charged ions derived from the linear C6H6 isomers. In the latter case Xe and benzene targets were used. The energetics and efficiency measurements for the former reactions suggest that for targets such as He and Ne the processes probably involve excited states of the doubly charged ions. The efficiencies measured for the latter reactions were distinctly different for the three C6H6 isomers and may indicate a strong dependence of charge exchange cross-section on doubly charged ion structure.  相似文献   

3.
The formation of carbon tetrachloride‐benzene charge transfer complex was confirmed by UV and NMR spectrometric studies. A change in UV spectrum of benzene is observed upon addition of carbon tetrachloride. Whereas the appearance of new bands supports the formation of charge transfer complex. NMR study shows that, chemical shift of benzene pmr signal depends on the CCl4‐C6H6 molar ratio. This observation is another criterion for the formation of benzene‐carbon tetrachloride charge transfer complex. Job's Continuous Variation method indicates that a 2:1 CCl4‐C6H6 charge transfer complex (2:1 CTC) is formed. The association constants (K2:1) of (2:1 CTC) was found to be 0.0197 M?2. The maximum concentration of (2:1 CTC) was found to be in samples with 2:1 CCl4‐C6H6 molar ratio (33% benzene mole). On the other hand the maximum yield of chlorobenzene was obtained, also, upon radiolysis of CCl4‐C6H6 samples at a 2:1 molar ratio (33% benzene mole). Therefore, it could be concluded that (2:1 CTC) participates in the formation of chlorobenzene upon radiolysis of the benzene‐carbon tetrachloride system. This conclusion was supported by the dependence of the chlorobenzene yield of a γ‐irradiated carbon tetrachloride‐benzene system (2:1 molar ratio) on irradiation time according to a third order kinetic equation with a very good linearity (R2 = 0.9977). Accordingly, the rate constant for the chlorobenzene formation under this condition was found to be ≈ 5.5 × 10?7 L2.mol?2.h?1. We propose a radiation chemical mechanism in which the 2:1 CTC plays a role in the formation of chlorobenzene.  相似文献   

4.
An essentially molecular ruthenium–benzene complex anchored at the aluminum sites of dealuminated zeolite Y was formed by treating a zeolite‐supported mononuclear ruthenium complex, [Ru(acac)(η2‐C2H4)2]+ (acac=acetylacetonate, C5H7O2?), with 13C6H6 at 413 K. IR, 13C NMR, and extended X‐ray absorption fine structure (EXAFS) spectra of the sample reveal the replacement of two ethene ligands and one acac ligand in the original complex with one 13C6H6 ligand and the formation of adsorbed protonated acac (Hacac). The EXAFS results indicate that the supported [Ru(η6‐C6H6)]2+ incorporates an oxygen atom of the support to balance the charge, being bonded to the zeolite through three Ru? O bonds. The supported ruthenium–benzene complex is analogous to complexes with polyoxometalate ligands, consistent with the high structural uniformity of the zeolite‐supported species, which led to good agreement between the spectra and calculations at the density functional theory level. The calculations show that the interaction of the zeolite with the Hacac formed on treatment of the original complex with 13C6H6 drives the reaction to form the ruthenium–benzene complex.  相似文献   

5.
The methyl viologen dication, used under the name Paraquat as an agricultural reagent, is a well‐known electron‐acceptor species that can participate in charge‐transfer (CT) interactions. The determination of the crystal structure of this species is important for accessing the CT interaction and CT‐based properties. The title hydrated salt, bis(1,1′‐dimethyl‐4,4′‐bipyridine‐1,1′‐diium) hexacyanidoferrate(II) octahydrate, (C12H14N2)2[Fe(CN)6]·8H2O or (MV)2[Fe(CN)6]·8H2O [MV2+ is the 1,1′‐dimethyl‐4,4′‐bipyridine‐1,1′‐diium (methyl viologen) dication], crystallizes in the space group P 21/c with one MV2+ cation, half of an [Fe(CN)6]4− anion and four water molecules in the asymmetric unit. The FeII atom of the [Fe(CN)6]4− anion lies on an inversion centre and has an octahedral coordination sphere defined by six cyanide ligands. The MV2+ cation is located on a general position and adopts a noncoplanar structure, with a dihedral angle of 40.32 (7)° between the planes of the pyridine rings. In the crystal, layers of electron‐donor [Fe(CN)6]4− anions and layers of electron‐acceptor MV2+ cations are formed and are stacked in an alternating manner parallel to the direction of the −2a + c axis, resulting in an alternate layered structure.  相似文献   

6.
Reactions of phenol and hydroxyl radical were studied under the aqueous environment to investigate the antioxidant characters of phenolic compounds. M06‐2X/6‐311 + G(d,p) calculations were carried out, where proton transfers via water molecules were examined carefully. Stepwise paths from phenol + OH + (H2O)n (n = 3, 7, and 12) to the phenoxyl radical (Ph O) via dihydroxycyclohexadienyl radicals (ipso, ortho, meta, and para OH‐adducts) were obtained. In those paths, the water dimer was computed to participate in the bond interchange along hydrogen bonds. The concerted path corresponding to the hydrogen atom transfer (HAT, apparently Ph OH + OH → Ph O + H2O) was found. In the path, the hydroxyl radical located on the ipso carbon undergoes the charge transfer to prompt the proton (not hydrogen) transfer. While the present new mechanism is similar to the sequential proton loss electron transfer (SPLET) one, the former is of the concerted character. Tautomerization reactions of ortho or para (OH)C6H5=O + (H2O)n → C6H4(OH)2(H2O)n were traced with n = 2, 3, 4, and 14. The n = 3 (and n = 14) model of ortho and para was calculated to be most likely by the strain‐less hydrogen‐bond circuit.  相似文献   

7.
The crystal structures of two proton‐transfer compounds of 3‐carb­oxy‐4‐hydroxy­benzene­sulfonic acid (5‐sulfosalicylic acid) with the aromatic polyamines 2,6‐diamino­pyridine [namely 2,6‐diamino­pyridinium 3‐carb­oxy‐4‐hydroxy­benzene­sulfonate monohydrate, C5H8N3+·C7H5O6S·H2O, (I)] and 1,4‐phenyl­ene­diamine [namely 1,4‐phenyl­ene­diaminium 3‐carboxyl­ato‐4‐hydroxy­benzene­sulfonate, C6H10N22+·C7H4O6S2−, (II)] have been determined. Both compounds feature extensively hydrogen‐bonded three‐dimensional layered polymer structures having significant inter­layer π–π inter­actions between the cation and anion species. In (I), the pyridine N atom of the Lewis base is protonated and forms a direct hydrogen‐bonding inter­action with the water mol­ecule, which together with the two amine groups of the cation and the carboxylic acid group of the anion also give additional inter­actions with O‐atom acceptors of the sulfonate group. In (II), a dianionic species results from deprotonation of both the sulfonic and the carboxylic acid groups, and all available O‐atom acceptors inter­act with all dication donors, which lie about inversion centres.  相似文献   

8.
The bimolecular reactivity of xenon with C7Hn2+ dications (n=6–8), generated by double ionization of toluene using both electrons and synchrotron radiation, is studied by means of a triple‐quadrupole mass spectrometer. Under these experimental conditions, the formation of the organoxenon dications C7H6Xe2+ and C7H7Xe2+ is observed to occur by termolecular collisional stabilization. Detailed experimental and theoretical studies show that the formation of C7H6Xe2++H2 from doubly ionized toluene (C7H82+) and xenon occurs as a slightly endothermic, direct substitution of dihydrogen by the rare gas with an expansion to a seven‐membered ring structure as the crucial step. For the most stable isomer of C7H6Xe2+, an adduct between the cycloheptatrienyldiene dication and xenon, the computed binding energy of 1.36 eV reaches the strength of (weak) covalent bonds. Accordingly, electrophiles derived from carbenes might be particularly promising candidates in the search for new rare‐gas compounds.  相似文献   

9.
One‐electron oxidation of the stibines Aryl3Sb ( 1 , Aryl=2,6‐i Pr2‐4‐OMe‐C6H2; 2 , Aryl=2,4,6‐i Pr3‐C6H2) with AgSbF6 and NaBArylF4 (ArylF=3,5‐(CF3)2C6H3) afforded the first structurally characterized examples of antimony‐centered radical cations 1 .+[BArylF4] and 2 .+[BArylF4]. Their molecular and electronic structures were investigated by single‐crystal X‐ray diffraction, electron paramagnetic resonance spectroscopy (EPR) and UV/Vis absorption spectroscopy, in conjunction with theoretical calculations. Moreover, their reactivity was investigated. The reaction of 2 .+[BArylF4] and p ‐benzoquinone afforded a dinuclear antimony dication salt 3 2+[BArylF4]2, which was characterized by NMR spectroscopy and X‐ray diffraction analysis. The formation of the dication 3 2+ further confirms that the isolated stibine radical cations are antimony‐centered.  相似文献   

10.
One‐electron oxidation of the stibines Aryl3Sb ( 1 , Aryl=2,6‐i Pr2‐4‐OMe‐C6H2; 2 , Aryl=2,4,6‐i Pr3‐C6H2) with AgSbF6 and NaBArylF4 (ArylF=3,5‐(CF3)2C6H3) afforded the first structurally characterized examples of antimony‐centered radical cations 1 .+[BArylF4] and 2 .+[BArylF4]. Their molecular and electronic structures were investigated by single‐crystal X‐ray diffraction, electron paramagnetic resonance spectroscopy (EPR) and UV/Vis absorption spectroscopy, in conjunction with theoretical calculations. Moreover, their reactivity was investigated. The reaction of 2 .+[BArylF4] and p ‐benzoquinone afforded a dinuclear antimony dication salt 3 2+[BArylF4]2, which was characterized by NMR spectroscopy and X‐ray diffraction analysis. The formation of the dication 3 2+ further confirms that the isolated stibine radical cations are antimony‐centered.  相似文献   

11.
OH addition reactions play a pivotal role in the atmospheric transformation of a number of phenyl and substituted phenyl‐based persistent and toxic organic pollutants. Here, we screened appropriate DFT functionals to predict reaction mechanisms and rate constants (kOH) of the OH additions by taking benzene and substituted benzenes (C6H5F, C6H5Cl, C6H5Br, C6H5CH3, C6H5OH) as model compounds. By comparing the kOH values calculated with DFT methods to experimental values, we found that the ωB97 functional is the best among the 18 functionals considered (using the basis sets 6‐31 + G(d,p) for optimizations and 6‐311++G(3df,2pd) for single point energy calculations) in the temperature range of 230‐330 K. In addition, we found that some other functionals performed well in specific conditions, e.g., BMKD3 is good for benzene, halogenated benzenes and C6H5CH3, and CAM‐B3LYP is good for the reaction of C6H5OH at room temperature. Based on the diversity of the electronic structures of the selected model compounds and the frequent occurrence of certain substituents ( CH3,  OH,  F,  Cl, and  Br) in the target compounds, the functionals recommended here can be used for future study of the reaction mechanisms and kOH values for OH addition to phenyl and substituted phenyl‐based persistent and toxic organic pollutants.  相似文献   

12.
A detailed computational study of the deamination reaction of melamine by OH, n H2O/OH, n H2O (where n = 1, 2, 3), and protonated melamine with H2O, has been carried out using density functional theory and ab initio calculations. All structures were optimized at M06/6‐31G(d) level of theory, as well as with the B3LYP functional with each of the basis sets: 6‐31G(d), 6‐31 + G(d), 6‐31G(2df,p), and 6‐311++G(3df,3pd). B3LYP, M06, and ω B97XD calculations with 6‐31 + G(d,p) have also been performed. All structures were optimized at B3LYP/6‐31 + G(d,p) level of theory for deamination simulations in an aqueous medium, using both the polarizable continuum solvation model and the solvation model based on solute electron density. Composite method calculations have been conducted at G4MP2 and CBS‐QB3. Fifteen different mechanistic pathways were explored. Most pathways consisted of two key steps: formation of a tetrahedral intermediate and in the final step, an intermediate that dissociates to products via a 1,3‐proton shift. The lowest overall activation energy, 111 kJ mol?1 at G4MP2, was obtained for the deamination of melamine with 3H2O/OH?.  相似文献   

13.
In the title 1/2/2 adduct, C4H12N22+·2C6H3N2O5?·2H2O, the dication lies on a crystallographic inversion centre and the asymmetric unit also has one anion and one water mol­ecule in general positions. The 2,4‐di­nitro­phenolate anions and the water mol­ecules are linked by two O—H?O and two C—H?O hydrogen bonds to form molecular ribbons, which extend along the b direction. The piperazine dication acts as a donor for bifurcated N—H?O hydrogen bonds with the phenolate O atom and with the O atom of the o‐nitro group. Six symmetry‐related molecular ribbons are linked to a piperazine dication by N—H?O and C—H?O hydrogen bonds.  相似文献   

14.
The reactions of methane with the dications C7H62+, C7H72+, and C7H82+ generated by electron ionization of toluene are studied using mass-spectrometry tools. It is shown that the reactivity is dominated by the formation of doubly charged intermediates, which can either eliminate molecular hydrogen to yield doubly charged products or undergo charge-separation reactions leading to the formation of a methyl cation and the corresponding C7Hn+1+ monocation. Typical processes observed for dications, like electron transfer or proton transfer, are largely suppressed. The theoretically derived mechanism of the reaction between C7H62+ and CH4 indicates that the formation of the doubly charged intermediate is kinetically preferred at low internal energies of the reactants. In agreement, the experimental results show a pronounced hydrogen scrambling and dominant formation of the doubly charged products at low collision energies, whereas direct hydride transfer prevails at larger collision energies.  相似文献   

15.
Structures of [C6H6]2+ ions formed by electron impact from benzene, 2,4-hexadiyne and 1,5-hexadiyne have been investigated by the recently developed electron capture induced decomposition method, by charge-separation reactions and by ion abundances in electron impact mass spectra. Significant differences were found among the isomers indicating the structural integrity of these [C6H6]2+ ions. The observed differences indicate that the most likely atomic configuration of [C6H6]2+ ions produced from benzene and 2,4-hexadiyne is the same as in the corresponding neutral species. A new method is suggested by which the structure (atomic configuration) of doubly charged ions may be determined.  相似文献   

16.
The title adduct, C5H14N22+·C8H3NO62−·C8H5NO6·H2O, crystallizes in the monoclinic space group P21. All O atoms of the 4‐nitro­phthalate anions and neutral 4‐nitro­phthalic acid mol­ecules are involved in hydrogen bonding with the piperazine dication and the water mol­ecule of crystallization.  相似文献   

17.
Reaction of the imidazolium-stabilized diphosphete-diide IDP with trityl phosphaalkyne affords a mixture which contains the molecules 1 a and 1 b with a central C3P3 core, which formally carries a two-fold negative charge. In order to avoid the formation of an antiaromatic 8π electron system within a conjugated dianionic six-membered [C3P3]2− ring, 1 a adopts a bicyclic [3.1.0] and 1 b a tricyclic [2.2.0.0] structure which are in a dynamic equilibrium. 1 a , b can be reversibly oxidized to a triphosphinine dication [ 5 ]2+ with a central flat aromatic six-membered C3P3 ring. This two-electron redox reaction occurs in two single-electron transfer steps via the 7π-radical cation [ 4 ]⋅+, which could also be isolated and fully characterized. The profound reversible structural change observed for the two-electron redox couple [ 5 ]2+/ 1 a , b is in sharp contrast to the C6H6/[C6H6]2− couple, which undergoes only a modest structural deformation.  相似文献   

18.
The reaction of fumaryl fluoride with the superacidic solutions XF/MF5 (X=H, D; M=As, Sb) results in the formation of the monoprotonated and diprotonated species, dependent on the stoichiometric ratio of the Lewis acid to fumaryl fluoride. The salts [C4H3F2O2]+[MF6] (M=As, Sb) and [C4H2X2F2O2]2+([MF6])2 (X=H, D; M=As, Sb) are the first examples with a protonated acyl fluoride moiety. They were characterized by low-temperature vibrational spectroscopy. Low-temperature NMR spectroscopy and single-crystal X-ray structure analyses were carried out for [C4H3F2O2]+[SbF6] as well as for [C4H4F2O2]2+([MF6])2 (M=As, Sb). The experimental results are discussed together with quantum chemical calculations of the cations [C4H4F2O2 ⋅ 2 HF]2+ and [C4H3F2O2 ⋅ HF]+ at the B3LYP/aug-cc-pVTZ level of theory. In addition, electrostatic potential (ESP) maps combined with natural population analysis (NPA) charges were calculated in order to investigate the electron distribution and the charge-related properties of the diprotonated species. The C−F bond lengths in the protonated dication are considerably reduced on account of the +R effect.  相似文献   

19.
Crystalline and properly ordered protonated benzene as the [C6H7]+[Al2Br7]??(C6H6) salt 1 are obtained by the combination of solid AlBr3, benzene, and HBr gas. Compound 1 was characterized and verified by NMR, Raman and X‐Ray spectroscopy. This unexpected simple and straight forward access shows that HBr/AlBr3 is an underestimated superacid that should be used more frequently.  相似文献   

20.
The electrospray ionization collisionally activated dissociation (CAD) mass spectra of protonated 2,4,6‐tris(benzylamino)‐1,3,5‐triazine (1) and 2,4,6‐tris(benzyloxy)‐1,3,5‐triazine (6) show abundant product ion of m/z 181 (C14H13+). The likely structure for C14H13+ is α‐[2‐methylphenyl]benzyl cation, indicating that one of the benzyl groups must migrate to another prior to dissociation of the protonated molecule. The collision energy is high for the ‘N’ analog (1) but low for the ‘O’ analog (6) indicating that the fragmentation processes of 1 requires high energy. The other major fragmentations are [M + H‐toluene]+ and [M + H‐benzene]+ for compounds 1 and 6, respectively. The protonated 2,4,6‐tris(4‐methylbenzylamino)‐1,3,5‐triazine (4) exhibits competitive eliminations of p‐xylene and 3,6‐dimethylenecyclohexa‐1,4‐diene. Moreover, protonated 2,4,6‐tris(1‐phenylethylamino)‐1,3,5‐triazine (5) dissociates via three successive losses of styrene. Density functional theory (DFT) calculations indicate that an ion/neutral complex (INC) between benzyl cation and the rest of the molecule is unstable, but the protonated molecules of 1 and 6 rearrange to an intermediate by the migration of a benzyl group to the ring ‘N’. Subsequent shift of a second benzyl group generates an INC for the protonated molecule of 1 and its product ions can be explained from this intermediate. The shift of a second benzyl group to the ring carbon of the first benzyl group followed by an H‐shift from ring carbon to ‘O’ generates the key intermediate for the formation of the ion of m/z 181 from the protonated molecule of 6. The proposed mechanisms are supported by high resolution mass spectrometry data, deuterium‐labeling and CAD experiments combined with DFT calculations. Copyright © 2012 John Wiley & Sons, Ltd.  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