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 15 毫秒
1.
Eleven isomers with the PyC2H 5 composition, which include three conventional (1–3) and eight distonic radical cations (4–11), have been generated and in most cases successfully characterized in the gas phase via tandem-in-space multiple-stage pentaquadrupole MS2 and MS3 experiments. The three conventional radical cations, that is, the ionized ethylpyridines C2H5-C5H4N (1–3), were generated via direct 70-eV electron ionization of the neutrals, whereas sequences of chemical ionization and collision-induced dissociation (CID) or mass-selected ion-molecule reactions were used to generate the distonic ions H2C·?C5H4N+?CH3 (4–6), CH3?C5H4N+?CH 2 · (7–9), C5H5N+?CH2CH 2 · (10), and C5H5N+?CH·?CH3 (11). Unique features of the low-energy (15-eV) CID and ion-molecule reaction chemistry with the diradical oxygen molecule of the isomers were used for their structural characterization. All the ion-molecule reaction products of a mass-selected ion, each associated with its corresponding CID fragments, were collected in a single three-dimensional mass spectrum. Ab initio calculations at the ROMP2/6–31G(d, p)//6–31G(d, p)+ZPE level of theory were performed to estimate the energetics involved in interconversions within the PyC2H5 system, which provided theoretical support for facile 4?7 interconversion evidenced in both CID and ion-molecule reaction experiments. The ab initio spin densities for the a-distonic ions 4–9 and 11 were found to be largely on the methylene or methyne formal radical sites, which thus ruled out substantial odd-spin derealization throughout the neighboring pyridine ring. However, only 8 and 9 (and 10) react extensively with oxygen by radical coupling, hence high spin densities on the radical site of the distonic ions do not necessarily lead to radical coupling reaction with oxygen. The very typical “spatially separated” ab initio charge and spin densities of 4–11 were used to classify them as distonic ions, whereas 1–3 show, as expected, “localized” electronic structures characteristic of conventional radical ions.  相似文献   

2.
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   

3.
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   

4.
Four novel calix[4]arene hydrazone-based receptors 3a?Cd were prepared in yields of 69?C87% by condensating formylated calix[4]arene ester (2) with salicylyl hydrazine, 2,4-dinitrophenyl hydrazine, nicotinyl hydrazine or phenyl thiosemicarhazide, respectively. New compounds were characterized through elemental analysis, IR, ESI?CMS, 1H NMR studies. Compounds 3a?Cd containing two binding sites had the complexation abilities for hard and soft cations concurrently. The noncompetitive extracting experiments showed compounds 3a?Cd were excellent receptors for hard and soft metal cations. The competitive extracting experiments exhibited the cooperative complexation in binding hard and soft metal cations and compound 3a possessed outstanding selectivity for Na+ and Hg2+. The IR spectra of compound 3a before and after complexation revealed that the soft metal cation was binded in the cavity composed of hydrazone groups and azo groups at the upper rims of calix[4]arene units and hard metal cations was binded in cavity composed of ester groups and phenolic hydroxyl groups at the lower rims of calix[4]arene units.  相似文献   

5.
The radical-molecule reaction of C2Cl3 with NO2 is explored at the B3LYP/6-311G(d,p) and CCSD(T)/6-311+G(d,p) (single-point) levels. On the singlet potential energy surface (PES), the association between C2Cl3 and NO2 is found to be carbon-to-nitrogen attack forming the adduct C2Cl3NO2 (1) without any encounter barrier, followed by isomerization to C2Cl3ONO (2). Starting from 2, the most feasible pathway is the N–O1 bond cleavage which lead to P 1 (C2Cl3O + NO). Much less competitively, 2 transforms to the three-membered ring isomer c-OCCl2C–ClNO (4 a ) which can easily interconvert to c-OCCl2C–ClNO 4 b . Then 4 (4 a , 4 b ) takes direct C1–C2 and C2–O1 bonds cleavage to give P 2 (COCl2 + ClCNO). The lesser competitive channel is the 4 a isomerizes to the four-membered ring intermediate O-c-CNClOCCl2 (5) followed by dissociation to P3 (CO + ClNOCCl2). The concerted 1,2-Cl shift along with C1–O1 bond rupture of 4 b to form ONC(O)CCl3 (6) followed by dissociation to P 4 (ClNO + OCCCl2) is even much less feasible. Moreover, some of P 3 and P 4 can further dissociate to P 5 (ClNO + CO + CCl2). Compared with the singlet pathways, the triplet pathways may have less contribution to the title reaction. Our results are in marked difference from previous theoretical studies which showed that two initial adducts C2Cl3–NO2 and C2Cl3–ONO are obtained. Moreover, in the present paper we focus our main attentions on the cyclic isomers in view of only the chain-like isomers are considered by previous studies. The present study may be helpful for understanding the halogenated vinyl chemistry.  相似文献   

6.
The condensation reactions of N2O3-donor type coronands (13) with hexachlorocyclotriphosphazatriene, N3P3Cl6, resulted in the formation of spiro-crypta phosphazene derivatives (46). These compounds with excess morpholine and 1,4-dioxa-8-azaspiro[4,5]decane (DASD) afford fully substituted morpholino (7 and 10) and 1,4-dioxa-8-azaspiro[4,5]deca (8)-substituted phosphazene derivatives, respectively. Whilst, in the same conditions, the reactions of 4, 5 and 6 with pyrrolidine, morpholine and DASD also produce partially pyrrolidino-substituted geminal (9 and 11), mono-substituted pyrrolidino (12), morpholino (13) and 1,4-dioxa-8-azaspiro[4,5]deca (14) phosphazenes. It has been clearly observed that the chloride replacement reactions of 4, 5 and 6 with pyrrolidine lead to the geminal products. Compounds 7, 8 and 10 are the first examples of anisochronic tetrakis (amino) phosphazenes according to 31P NMR data. The structures of 7, 8 and 1014 have been determined by FTIR, MS, 1H, 13C and 31P NMR, DEPT, and HETCOR spectral data. The solid-state structures of 9, 13 and 14 have been examined by X-ray diffraction techniques. The sums of the bond angles around the spiro cyclic nitrogen atoms [344.8(4)° and 347.6(4)°] of 9, indicate that the nitrogen atoms have pyramidal geometries. Thus, the N atoms seem to have stereogenic configurations. Compounds 1214 also have two stereogenic P-atoms, and they are expected to be in the mixture of enantiomers. The relationships between NPN (α and α′) bond angles and δPspiro values and the correlation of Δ(P–N) with δPspiro and Δ(δP) values are presented.  相似文献   

7.
The unimolecular dissociation of (CH3)2C+OC2H5 ions (I) and their deuterated analogs, generated by ion-molecule reactions (IMR) in acetone-ethyl iodide mixtures was studied by tandem mass Spectrometry methods. Two significant processes that yielded I ions were identified. The Fourier transform ion cyclotron resonance study showed that the reaction between ionized ethyl iodide and neutral acetone was the principal source of I. This process involved the formation of the stable mixed ionized dimer, [C2H5I·O=C(CH3)2] (II), which dissociated by the loss of an I atom. Other important fragmentation pathways of II were the formation of C2H5I, (CH3)2CO; and (CH3)2COI+ and the loss of CH3CHI·. The major dissociation of I was the loss of C2H4. The activation energy for this reaction was determined by metastable ion appearance energy measurements to be ~55 kJ mol?1 above the thermochemical minimum. The analysis of the metastable and collision-induced dissociation of D-labeled I showed an unusual time-energy effect on the degree of H/D mixing, with the highest selectivity for the ethene loss [β-H(D)-atom shift] being observed for ions with the lowest internal energies. Collisional excitation could not produce significant H/D mixing among dissociating ions. The results were rationalized by the existence of two species— the classical (2-ethoxypropyl) and nonclassical (proton-bound acetone-ethene pair) isomers of I. The classical structure was originally formed by IMR or from II. The energy barrier for the classical to nonclassical isomerization lay well above the thermochemical threshold for C2H4 loss, providing only limited H-atom mixing in nonclassical ions that were always formed in their dissociative state. The effect of the proton affinity of the carbonyl compound on the H/D mixing in RR′C+OC2H5 ions was studied. It was shown that the selectivity for the ethene loss (β-H-atom shift) generally increased with the increase of the proton affinity of RR′CO. Neutralization-reionization mass spectrometry was applied to a study of (CH3)2C+OR ions, where R = H, I, C2H5. The observation of a recovery signal for the ion I was attributed to the formation of the 2-ethoxypropyl radical. Neutral counterparts of (CH3)2COI+ ions were also generated, being the first example of IO-substituted alkyl radicals.  相似文献   

8.
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   

9.
The repulsive nature of the interaction between the cation radicals of the π-[terthiophene] 2 2+ dimers, 1 2 2+ , found in crystals has been concluded from B3LYP/6-31+G(d) calculations. Hence, the bonding component is weaker than the Coulombic repulsion, consistent to recent findings for [TTF]·+–[TTF]·+ interactions (TTF = tetrathiafulvalene). The existence of 1 2 2+ dimers originates from the cation+–anion? electrostatic interactions, which exceeds the combined effect of the 1 .+1 .+ plus (SbF6)?–(SbF6)? repulsions in 1 2(SbF6)2, similar to what is found for [TTF]·+–[TTF]·+ interactions in [TTF]2(ClO4)2 aggregates and in crystals. The long, multicenter bond in 1 2 2+ is characterized as a 2e?/10c bond from an Atoms-in-molecules analysis.  相似文献   

10.
Two different hydrogen-bonded inclusion compounds, [2,4,6-C5H2N(COO?)3]0.5·[C(NH2) 3 + ]0.5·[(C2H5)4N+]·2H2O (1) and [2,4,6-C5H2N(COO?)3]·[C(NH2) 3 + ]·[(C2H5)4N+]·[(C3H7)4N+]·6H2O (2) are reported in this paper, in which 2,4,6-pyridine-tricarboxylic anions, guanidiniums and water molecules jointly construct host lattices while tetraalkylammonium cations are accommodated as guest species. Both two compounds formed sandwich-like hydrogen-bond inclusion compounds. In compound 1, the dimers composed of 2,4,6-pyridine-tricarboxylic anions and guanidiniums form 2D hydrogen-bonded layers by connecting with water molecules. In compound 2, 2,4,6-pyridine-tricarboxylic anions, guanidiniums and water molecules contribute to generate an undulate rosette hydrogen-bonded architecture. Interestingly, in compound 2, there are two species of guest molecules, tetraethylammonium and tetrapropylammonium, which are alternately arranged between the neighboring layers. Mixed guest cations accommodated in hydrogen-bonded inclusion compounds are seldom seen.  相似文献   

11.
Both the singlet and triplet potential energy surfaces (PESs) of the NH (X3Σ?) + HCNO reaction have been investigated at the BMC-CCSD level based on the UB3LYP/6-311++G(d, p) structures. The results show that the title reaction is more favorable through the singlet potential energy surface than the triplet one. For the singlet potential energy surface of the NH (X3Σ?) + HCNO reaction, the most feasible association of NH (X3Σ?) with HCNO is found to be a non-barrier nitrogen-to-carbon attack forming the adduct a (trans-HNCHNO), which can isomerize to the adduct b (cis-HNCHNO). The most feasible channel is that the 1, 3-H shift with N2–H2 and C–N1 bonds cleavage associated with the N1–H2 bond formation of adduct a leads to the product P 1 (HCN + HNO). Moreover, P 2 (HNC + HNO) should be the competitive product. The other products, including P 3 (NH2 + NCO) and P 4 (N2H2 + CO), are minor products. The product P 1 can be obtained through two competitive channels Path 1: R  a  P 1 and Path 3: R  b  d  P 1 , whereas the product P 2 can be formed through Path 2: R  b  d  P 2 . At high temperatures, the nitrogen-to-nitrogen approach may become feasible. For the triplet potential energy surface of the NH (X3Σ?) + HCNO reaction, the Path 10: R  3 a  3 a 1  P 1 should be the most feasible pathway due to the less reaction steps and lower barriers. These conclusions will have impacts on further experimental investigations.  相似文献   

12.
Imidazole base was crystallized with different aromatic carboxylic acids 2,4-dihydroxybenzoic acid, 5-chlorosalicylic acid, and 1,8-naphthalic acid, affording three new binary molecular organic salts of [(C 3 H 5 N 2 + )·(C 7 H 5 O 4 )] (1), [(C 3 H 5 N 2 + )·(C 7 H 4 O 3 Cl )] C 7 H 5 O 3 Cl (2), and [(C 3 H 5 N 2 + ) (C 12 H 7 O 4 )] (3). Proton transfer occurs from the COOH of carboxylic acid to nitrogen of imidazole in all complexes (1-3), leading to the hydrogen bond N-H…O in all structures. To our knowledge, the recognition pattern between the carboxylic acid group and imidazole (acid-imidazole synthon) is less well-studied so far. The cooperation among COOH, COO and imidazolium cation functional groups for the observed hydrogen bond synthons is examined in the three structures. Generally, the strong N-H…O and O-H…O hydrogen bonds define supramolecular architecture and connectivity within chains, while weaker C-H…O hydrogen bonds play the dominant role in controlling the interactions between layers in these novel organic salts. Thermal stability of these compounds has been investigated by thermogravimetric analysis (TGA) of mass loss.  相似文献   

13.
Energetic, geometric and magnetic criteria were applied to examine the stability and/or aromatic character for the cyclic molecules C 4 H 4 M (M = O, S, Se, Te, NH, PH, AsH and SbH) at B3LYP/6-311++G** and MP2/6-311++G** levels of theory. The isodesmic reactions and nuclear independent chemical shifts (NICS) calculations were utilized to examine the molecules for energetic and magnetic criteria, respectively. The isodesmic reaction energies reveal that thiophene (C 4 H 4 S, ?23.269 kcal/mol) and pyrrole (C 4 H 4 NH, ?20.804 kcal/mol) have the greatest aromatic stabilization energies and tellurophene (C 4 H 4 Te, ?15.114 kcal/mol) and stibole (C 4 H 4 SbH, ?1.169 kcal/mol) have the lowest aromatic stabilization energies in their corresponding groups at MP2/6-311++G**. The NICS calculations confirmed the results obtained through isodesmic reaction energies.  相似文献   

14.
The complex triplet potential energy surface for the reaction of HCNO with NH is investigated at the G3B3 level using the B3LYP/6-311++G(d,p), and QCISD/6-311++G(d,p) geometries. Various possible isomerization and dissociation pathways are probed. The initial association between HCNO and NH is found to be carbon to nitrogen attack leading to HNCHNO 2a, which can convert to 2b, 2c, and 2d. Subsequently, 1,4-H-shift of 2a to form NCHNOH 3a followed by dissociation to P 2 (1HCN + 3HON) is the most feasible pathway. Much less competitively, 2d undergoes successive 1,3-H-shift and C-N cleavage to form HNCNOH 8b, and then to product P 3 (1HNC + 3HON), the second feasible pathway. 8b can alternatively isomerize to 8c followed by N–O bond rupture to generate P 6 (2OH + 2HNCN), the lesser followed feasible pathway. In addition, 2b takes continuously 1,3- and 1,2-H-shift to form NC(H)NHO 6a, then to ONHCNH 7a which can convert to 7b. Eventually, 7b may take C-N bond fission to produce P 5 (1HNC + 3HNO), the least feasible pathway. The present paper may be helpful for future experimental identification of the product distributions for the title reaction, and may be helpful to deeply understand the mechanism of the title reaction.  相似文献   

15.
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   

16.
Isomeric structures, energies, and properties of silacyclopropylidenoids, C2H4SiMX (where M?=?Li or Na and X?=?F, Cl or Br), were studied ab initio at the HF and MP2 levels of theory using the 6-31+G(d,p) and aug-cc-pVTZ basis sets. The calculations indicate that each of C2H4SiMXs has three stationary structures: silacyclopropylidenoid (S), tetrahedral (T), and inverted (I). All of the silacyclopropylidenoid (S) forms are energetically more stable than others except that S-LiF is by only 0.7?kcal/mol higher in energy than I-LiF. In contrast, all of the tetrahedral (T) forms are the most unstable ones except for T-NaF. Energy differences between S, T, and I forms range from 0.70 to 8.70?kcal?mol?1 at the MP2/6-31+G(d,p) level. In addition, the molecular electrostatic potential maps, natural bond orbitals, and frontier molecular orbitals were calculated at the MP2/6-31+G(d,p) level.  相似文献   

17.
Metastable ion decompositions, collision-activated dissociation (CAD), and neutralization-reionization mass spectrometry are utilized to study the unimolecular chemistry of distonic ion ·CH2CH2CH?OH (2) and its enol-keto tautomers CH3CH=CHOH (1 ) and CH3CH2CH=O (3). The major fragmentation of metastable 1–3 is H· loss to yield the propanoyl cation, CH3CH2C≡O+. This reaction remains dominant upon collisional activation, although now some isomeric CH2=CH-CH+ OH is coproduced from all three precursors. The CAD and neutralization-reionization (+NR+) spectra of keto ion 3 are substantially different from those of tautomers 2 and 1. Hence, 3 without sufficient energy for decomposition (i. e. , “stable” 3) does not isomerize to the ther-modynamically more stable ions 2 or 1, and the 1,4-H rearrangement H-CH2CH2CH=O(3 ) → CH2CH2CH+ O-H (2 ) must require an appreciable critical energy. Although the fragment ion abundances in the + NR + (and CAD) spectra of 1 and 2 are similar, the relative and absolute intensities of the survivor ions (recovered C3H6O ions in the +NR+ spectra) are markedly distinct and independent of the internal energy of 1 and 2 . Furthermore, 1 and 2 show different MI spectra. Based on these data, distonic ion 2 does not spontaneously rearrange to enol ion 1 (which is the most stable C3H6O of CCCO connectivity) and, therefore, is separated from it by an appreciable barrier. In contrast, the molecular ions of cyclopropanol (4 ) and allyl alcohol (5 ) isomerize readily to 2 , via ring opening and 1,2-H? shift, respectively. The sample found to generate the purest 2 is α-hydroxy-γ-butyrolactone. Several other precursors that would yield 2 by a least-motion reaction cogenerate detectable quantities of enol ion 1 , or the enol ion of acetone (CH2=C(CH3)OH, 6 ), or methyl vinyl ether ion (CH3OCH=CH 2 , 7 ). Ion 6 is coproduced from samples that contain the —CH2—CH(OH)—CH2— substructure, whereas 7 is coproduced from compounds with methoxy substituents. Compared to CAD, metastable ion characteristics combined with neutralization-reionization allow for a superior differentiation of the ions studied.  相似文献   

18.
The zirconium nitrate complexes (NO2)[Zr(NO3)3(H2O)3]2(NO3)3 (1), Cs[Zr(NO3)5] ((2), (NH4)[Zr(NO3)5](HNO3) (3), and (NO2)0.23(NO)0.77[Zr(NO3)5] ((4) were prepared by crystallization from nitric acid solutions in the presence of H2SO4 or P2O5. The complexes were characterized by X-ray diffraction. The crystal structure of 1 consists of nitrate anions, nitronium cations, and [Zr(NO3)3(H2O)3]+ complex cations in which the ZrIV atom is coordinated by three water molecules and three bidentate nitrate groups. The coordination polyhedron of the ZrIV atom is a tricapped trigonal prism formed by nine oxygen atoms. The island structures of 2 and 3 contain [Zr(NO3)5]? anions and Cs+ or NH4 + cations, respectively. In addition, complex 3 contains HNO3 molecules. Complex 4 differs from (NO2)[Zr(NO3)5] in that three-fourth of the nitronium cations in 4 are replaced by nitrosonium cations NO+, resulting in a decrease in the unit cell parameters. In the [Zr(NO3)5]? anion involved in complexes 2–4, the ZrIV atom is coordinated by five bidentate nitrate groups and has an unusually high coordination number of 10. The coordination polyhedron is a bicapped square antiprism.  相似文献   

19.
The acyclic iminate amine gold(III) complex iodide [Au(C9H19N4)]I2 (I) was obtained. The crystal structure of complex I is stabilized by weak secondary interactions N-H??I, C-H??Y (Y = I, Au, and ??), and Au??I forming an infinite 2D network. Double stacks of cations along the axis x are united through zigzag chains of H-bonds and of secondary contacts. Structure I contains 3?C14-membered cation-cation and cation-anion rings. The anions I(1)? and I(2)? act as bridges in structure I.  相似文献   

20.
One-electron oxidation of 2-alkyl-1,4-dimethoxybenzenes 1a-f (2-alkyl=Me, Et, i-Pr, cy-C3H5CH2, PhCH2 and t-Bu) by 4-nitrobenzoyl peroxide 2 and pentaflurobenzoyl peroxide 3 was proved by the observation of great acceleration of decomposition of the peroxides at room temperature, the detection of the corresponding radical cations 1 +? a-f and product analysis. The product studies have disclosed that under the conditions employed (in acetonitrile at 40°C), the reaction pathways of the radical cations are greatly dependent on the nature of 2-alkyl substituents: Ring-4-nitrobenzoloxylation product at C 5 and C 6 were obtained exclusively in the reactions of the donors with aliphatic 2-alkyl substituents bearing at least one α-hydrogen atom, such as 1a, 1b, 1c and 1d; whereas in the case of 1e (with 2-benzyl group), both ring-substitution at C 5 (4e) and C 6 (5e) and deprotonation/4-nitrobenzoloxylation products 8e were isolated; from the donor without α-hydrogen atom, 1f, de-t-butylation products 12 and t-butyl 4-nitrobenzoate 13 were incorporated with ring-substitution at C 5 (4f) and C 6 (5f). Furthermore, the product distribution (4 over 5) is also affected by the bulkiness of 2-alkyl group. For all the electron-transfer reactions, large amounts of the benzoic acid (4-NO2-C6H4COOH or C6F5COOH) were generated and trace amounts of de-methylation product (2-alkyl-1,4-benzoqinones 6) were also detected by 1H NMR.  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