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 62 毫秒
1.
The reaction of the nitrone spin trap 5,5‐dimethylpyrroline‐N‐oxide (DMPO) with sodium (bi)sulfite in aqueous solutions was investigated using NMR and EPR techniques. Reversible nucleophilic addition of (bi)sulfite anions to the double bond of DMPO was observed, resulting in the formation of the hydroxylamine derivative 1‐hydroxy‐5,5‐dimethylpyrrolidine‐2‐sulfonic acid, with characteristic 1H and 13C NMR spectra. The reaction mechanism was suggested and corresponding equilibrium constants determined. The mild oxidation of the hydroxylamine results in the formation of an EPR‐detected spectrum identical with that for the DMPO adduct with sulfur trioxide anion radical. The latter demonstrates that a non‐radical addition reaction of (bi)sulfite with DMPO may contribute to the EPR detection of SO3?? radical. This possibility must be taken into account in spin trapping analysis of sulfite radical. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

2.
Contact with SO2 causes almost immediate dissolution of tetraalkylammonium halides, R4NX, (R = CH3 (Me), X = I; R = C2H5 (Et), X = Cl, Br, I; R = C4H9 (nBu), X = Cl, Br), with the formation of an adduct, [R4N]+[(SO2)nX] (n = 1–4). Vapor pressure measurements indicate the proclivity for SO2 uptake follows the order N(CH3)4+ < N(C2H5)4+ < N(C4H9)4+. This trend is in accord with the Jenkins–Passmore volume‐based thermodynamic model. Born–Haber cycles, incorporating the lattice energy and gas phase energy terms, are used to evaluate the energetic feasibility of reactions. Density functional theory calculations (B3PW91; 6‐311+G(3df)) have been used to calculate the energetics of (SO2)nX (X = Cl and Br) anions in the gas phase. The experimental studies show that tetraalkylammonium halides are feasible sorbents for SO2. In order to correlate the theoretical model, experimental enthalpy, Δr and entropy, Δr changes have been determined by the van't Hoff method for the binding of one SO2 molecule to (C2H5)4NCl, resulting in the liquid adduct (C2H5)4NCl · SO2. The structure of the analogous 1:1 bromide adduct, (C2H5)4NBr · SO2, has been determined by single‐crystal X‐ray diffraction (monoclinic, P21/c, a = 9.1409(14) Å, b = 12.3790(19) Å, c = 11.3851(17) Å, β = 107.952(2)°, V = 1225.6(3) Å3). The structure consists of discrete alkylammonium cations, bromide anions and SO2 molecules with short contacts between the anion and SO2 molecules. The (C2H5)4N+ cationadopts a transoid conformation with D2d symmetry, and represents a rare example of a well‐ordered (C2H5)4N+ cation in a crystal structure. The Br anions and SO2 molecules forms a chain, (SO2Br)n, with bifurcated contacts. Non‐bonding electron pairs on the halide anions engage in electrostatic interactions with the sulfur atoms and charge‐transfer interactions with the antibonding S–O orbitals of the bound SO2 moiety. Raman and 17O NMR spectra provide compelling evidence for a charge‐transfer interaction between SO2 molecules and the halide ions.  相似文献   

3.
The possible stable forms and molecular structures of 1‐cyclohexylpiperazine (1‐chpp) and 1‐(4‐pyridyl)piperazine (1‐4pypp) molecules have been studied experimentally and theoretically using nuclear magnetic resonance(NMR) spectroscopy. 13C, 15N cross‐polarization magic‐angle spinning NMR and liquid phase1H, 13C, DEPT, COSY, HETCOR and INADEQUATE NMR spectra of 1‐chpp (C10H20N2) and 1‐4pypp (C9H13N2) have been reported. Solvent effects on nuclear magnetic shielding tensors have been investigated using CDCl3, CD3 OD, dimethylsulfoxide (DMSO)‐d6, (CD3)2CO, D2O and CD2Cl2. 1H and 13C NMR chemical shifts have been calculated for the most stable two conformers, equatorial–equatorial (e–e) and axial–equatorial (a–e) forms of 1‐chpp and 1‐4pypp using B3LYP/6‐311++G(d,p)//6‐31G(d) level of theory. Results from experimental and theoretical data showed that the molecular geometry and the mole fractions of stable conformers of both molecules are solvent dependent. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

4.
Preparation of Tetramethylammonium Azidosulfite and Tetramethylammonium Cyanate Sulfur Dioxide‐Adduct, [(CH3)4N]+[SO2N3], [(CH3)4N]+[SO2OCN] and Crystal Structure of [(CH3)4N]+[SO2N3] Tetramethylammonium azide forms with sulfur dioxide an azidosulfite salt. It is characterized by NMR and vibrational spectroscopy and the crystal structure analysis. [(CH3)4N]+[SO2N3] crystallizes in the monoclinic space group P21/c with a = 551.3(1) pm, b = 1095.2(1) pm, c = 1465.0(1) pm, β = 100.63(1)°, and four formula units in the unit cell. The crystal structure possesses a strong S–N interaction between the N3– anions and the SO2 molecules. The S–N distance of 200.5(2) pm is longer than a covalent single S–N bond. The structure is compared with ab initio calculated data. Furthermore an adduct of tetrametylammonium cyanate and sulfur dioxide is reported. It is characterised by NMR and vibrational spectroscopy. The structure is calculated by ab initio methods.  相似文献   

5.
Five mono‐nuclear silver(I) complexes with the ligand 2,9‐dimethyl‐1,10‐phenanthroline, namely [Ag(DPEphos)(dmp)]BF4 ( 1 ), [Ag(DPEphos)(dmp)]CF3SO3 ( 2 ), [Ag(DPEphos)(dmp)]ClO4 ( 3 ), [Ag(DPEphos)(dmp)]NO3 ( 4 ), and [Ag(dppb)(dmp)]NO3 · CH3OH ( 5 ) {DPEphos = bis[2‐(diphenylphosphanyl)phenyl]ether, dppb = 1,2‐bis(diphenylphosphanyl)benzene, dmp = 2,9‐dimethyl‐1,10‐phenanthroline} were characterized by X‐ray diffraction, IR, 1H NMR, 31P NMR and fluorescence spectroscopy. Their terahertz (THz) time‐domain spectra were also studied. In these complexes the silver(I), which is coordinated by two kinds of chelating ligands, adopts four‐coordinate modes to generate mono‐nuclear structures. In complexes 1 , 3 – 5 , offset π ··· π weak interactions exist between the neighboring benzene rings. In the 31P NMR spectra, there exist splitting signals (dd), which can be attributed to the coupling of the 107,109Ag–31P. All the emission peaks of these complexes are attributed to ligand‐centered excited states.  相似文献   

6.
Copolymers of ethylene and sulfur dioxide containing 40–60 wt-% sulfur dioxide have been analyzed by using 220 MHz high-resolution NMR, and it has been shown that they contain structures of the form, ? SO2? (CH2? CH2)n? SO2? , where n is 1, 2, 3, 4, … The relative numbers of structures with n = 1, 2, 3, or 4 and above can be calculated from the NMR spectra. The fraction of ethylenes in longer blocks and the sulfur dioxide contents of the polymers can also be determined from the NMR data. The NMR results indicate that the distribution of ethylenes among the different structures is not that expected for a random copolymerization of ethylene and sulfur dioxide but that the arrangement of these structures within the copolymer is random.  相似文献   

7.
A series of aryl‐substituted enaminoketones and their thio analogues in CDCl3 solution and in the solid state were studied by the use of high‐resolution 1H and 13C as well as 13C cross polarization magic angle spinning (CP MAS) NMR spectra in combination with gauge including atomic orbitals‐density functional theory (GIAO‐DFT) calculations performed at the B3PW91/6–311 + + G(d,p) level of theory using the B3PW91/6‐311 + + G(d,p)‐optimized geometries. The analysis of the 13C NMR spectra in solution was done by using the Incredible Natural Abundance DoublE QUAntum Transfer Experiment (INADEQUATE) technique, whereas trends observed in the 13C shielding constants, calculated for the compounds studied, were a great help in assigning most of the signals in the 13C CP MAS NMR spectra. It was established on the basis of the experimental and theoretical NMR data that both groups of compounds exist in the form of Z‐s‐Z‐s‐E isomers in CDCl3 solution as well as in the solid state, with the NH hydrogen atom involved in intramolecular hydrogen bonding. This conclusion is in agreement with the fact that some of the compounds studied reveal liquid‐crystalline properties. Three‐bond H, H and C, H coupling constants measured in solution played a crucial role in the structure elucidation. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

8.
The 33S NMR signal of gaseous carbonyl sulfide (COS) was monitored as a function of density for the first time. An extrapolation to the zero‐density limit permitted the measurement of nuclear magnetic shielding of an isolated COS molecule. An improved 33S shielding scale was established taking the value of 817(12) ppm as the absolute shielding of COS. The new 33S shielding scale is certainly more accurate than any previous estimation and contains some reference standards, e.g. an isolated SF6 molecule, a saturated solution of (NH4)2SO4 in D2O, 2 M aqueous Cs2SO4 solution or liquid SF6, CS2 and SO2. The latter results can be applied for the easy estimation of sulfur shielding available from all the measurements of 33S NMR chemical shifts. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

9.
The effect of CD-inclusion on spin-trapping rates and spin-adduct decay rates for sulfur trioxide radical anion (SO3 ??) was investigated. SO3 ?? radical was produced with UV photolysis of sodium sulfite in basic aqueous solution, and spin-trapped with various spin traps, i.e., PBN (α-phenyl-N-t-butylnitrone), DMPO (5,5-dimethyl pyrroline-1-oxide), and three other phosphoryl DMPO-type spin traps. A modified β-CD, 6-O-α-d-glucosyl-β-cyclodextrin (G-β-CD) having better inclusion properties than β-CD, was employed. Upon adding excess G-β-CD, decay rates of SO3 ?? radical adducts significantly decreased in most spin traps. Half-lives of SO3 ?? radical adducts of phosphoryl spin traps were one to two orders of magnitude longer than that of PBN or DMPO, and the G-β-CD addition further extended the half-life time. The spin traps containing phosphoryl-group all showed higher SO3 ?? trapping rates than those of PBN and DMPO, but two phosphoryl spin traps achieved slower trapping rates by G-β-CD addition. In addition, the structures of CD-inclusion complexes of spin traps were established by means of 1D and 2D NMR measurements. Based on the results, the influences of inclusion on the spin-trapping rate processes and spin-adduct stabilizations were discussed. We conclude that substituents in DMPO-type spin traps may be modified to provide best spin-trapping capabilities in the presence or absence of CD.  相似文献   

10.
Silver triflate [AgOTf] assisted de‐bromination gives [Ni(dppm/dppe/(PPh3)2) (OTf)2], which on reaction with 4,4′‐bpy and gold(I) phosphines in dichloromethane medium by the self assemble technique leads to [{(L)Ni}{(4,4‐bpy)Au(PPh3)}2](OTf)4, ( 1,2,3 ) [{(L)Ni(4,4‐bpy)}4](OTf)8, ( 4,5,6 ) [L = dppm/dppe/(PPh3)2 = diphenyl phosphino‐methane, ‐ethane, bis‐triphenylphosphine, OSO2CF3 is the triflate anion]. The maximum molecular peak of the corresponding molecule is observed in the ESI mass spectrum. Ir spectra of the complexes show ‐C=C‐, ‐C=N‐, as well as phosphine stretching. The 1H NMR spectra as well as 31P (1H)NMR suggest solution stereochemistry, proton movement, and phosphorus proton interaction. Considering all the moieties, there are a lot of carbon atoms in the molecule reflected by the 13C NMR spectrum. In the 1H‐1H COSY spectrum of the present complexes and contour peaks in the 1H?13C HMQC spectrum, we assign the solution structure and stereoretentive transformation in each step.  相似文献   

11.
Detailed solution‐NMR studies on the distorted ruthenium hydride complex [RuH(η6‐toluene)(Binap)](CF3SO3) (2) are reported. NOE‐spectroscopy, together with low‐temperature 1H and 31P NMR data, reveals restricted rotation around a P—C bond for a specific axial P—phenyl ring with the activation energy determined via simulation. From 19F, 1H HOESY data, the approach of the triflate anion relative to the hydride ligand is established. Comparison of the quadrupole coupling constant CQF from both solution‐ and solid‐state MAS‐NMR on the deuteride [RuD(η6‐benzene)(Binap)](CF3SO3) (1‐D) provide information on the nature of the Ru—H bond. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

12.
In the negative‐ion collision‐induced dissociation mass spectra of most organic sulfonates, the base peak is observed at m/z 80 for the sulfur trioxide radical anion (SO3–·). In contrast, the product‐ion spectra of a few sulfonates, such as cysteic acid, aminomethanesulfonate, and 2‐phenylethanesulfonate, show the base peak at m/z 81 for the bisulfite anion (HSO3). An investigation with an extensive variety of sulfonates revealed that the presence of a hydrogen atom at the β‐position relative to the sulfur atom is a prerequisite for the formation of the bisulfite anion. The formation of HSO3 is highly favored when the atom at the β‐position is nitrogen, or the leaving neutral species is a highly conjugated molecule such as styrene or acrylic acid. Deuterium‐exchange experiments with aminomethanesulfonate demonstrated that the hydrogen for HSO3 formation is transferred from the β‐position. The presence of a peak at m/z 80 in the spectrum of 2‐sulfoacetic acid, in contrast to a peak at m/z 81 in that of 3‐sulfopropanoic acid, corroborated the proposed hydrogen transfer mechanism. For diacidic compounds, such as 4‐sulfobutanoic acid and cysteic acid, the m/z 81 ion can be formed by an alternative mechanism, in which the negative charge of the carboxylate moiety attacks the α‐carbon relative to the sulfur atom. Experiments conducted with deuterium‐exchanged and deuterium‐labeled analogs of sulfocarboxylic acids demonstrated that the formation of the bisulfite anion resulted either from a hydrogen transfer from the β‐carbon, or from a direct attack by the carboxylate moiety on the α‐carbon. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

13.
Lanthanum‐139 NMR spectra of stationary samples of several solid LaIII coordination compounds have been obtained at applied magnetic fields of 11.75 and 17.60 T. The breadth and shape of the 139La NMR spectra of the central transition are dominated by the interaction between the 139La nuclear quadrupole moment and the electric field gradient (EFG) at that nucleus; however, the influence of chemical‐shift anisotropy on the NMR spectra is non‐negligible for the majority of the compounds investigated. Analysis of the experimental NMR spectra reveals that the 139La quadrupolar coupling constants (CQ) range from 10.0 to 35.6 MHz, the spans of the chemical‐shift tensor (Ω) range from 50 to 260 ppm, and the isotropic chemical shifts (δiso) range from ?80 to 178 ppm. In general, there is a correlation between the magnitudes of CQ and Ω, and δiso is shown to depend on the La coordination number. Magnetic‐shielding tensors, calculated by using relativistic zeroth‐order regular approximation density functional theory (ZORA‐DFT) and incorporating scalar only or scalar plus spin–orbit relativistic effects, qualitatively reproduce the experimental chemical‐shift tensors. In general, the inclusion of spin–orbit coupling yields results that are in better agreement with those from the experiment. The magnetic‐shielding calculations and experimentally determined Euler angles can be used to predict the orientation of the chemical‐shift and EFG tensors in the molecular frame. This study demonstrates that solid‐state 139La NMR spectroscopy is a useful characterization method and can provide insight into the molecular structure of lanthanum coordination compounds.  相似文献   

14.
A new two‐step route toward the synthesis of polymeric ionic liquid microgel particles is presented. In the first step, hydrophilic microparticles were prepared by the concentrated emulsion polymerization of the ionic liquid 1‐vinyl‐3‐ethylimidazolium bromide in the presence of small amounts of N,N‐dimethylenebisacrylamide as a crosslinking agent. In the second step, the bromide anion was exchanged in water with different anions such as BF, CF3SO, (CF3SO2)2N?, (CF3CF2SO2)2N?, and dodecylbenzenesulfonate, and this resulted in the coagulation of the microparticles, which were easily recovered by filtration. The obtained polymeric ionic liquid microparticles could be swollen in a very broad range of organic solvents, including apolar organic solvents. As an application, glucose oxidase was encapsulated inside polymeric ionic liquid microparticles, which were used in an amperometric biosensor. The response of the biosensor showed excellent values that strongly depended on the nature of the polymeric ionic liquid counteranion in the order of Br? > BF > (CF3SO2)2N?. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 3958–3965, 2006  相似文献   

15.
Potassium oxosulfatovanadate(V) K3VO2(SO4)2 has been obtained by solid-phase synthesis from K2SO4, K2S2O7, and V2O5 (2: 1: 1), and its formation conditions, crystal structure, and physiochemical properties have been studied. The conversions of K3VO2(SO4)2 in contact with potassium vanadates and other potassium oxosulfatovanadates(V) are considered in terms of phase relations in the K2O-V2O5-SO3 system, which models the active component of vanadium catalysts for sulfur dioxide oxidation into sulfur trioxide. The X-ray diffraction pattern of K3VO2(SO4)2 is indexed in the monoclinic system (space group P21) with unit cell parameters of a = 10.0408(1) Å, b = 7.2312(1) Å, c = 7.3821(1) Å, β = 104.457(1)°, Z = 2, and V = 519.02 Å3. The crystal structure of K3VO2(SO4)2 is built from [VO2(SO4)2]3? complex anions, in which the vanadium atom is in an octahedral oxygen environment formed by two terminal oxygen atoms (V-O(6) = 1.605(7) Å, V-O(10) = 1.619(7) Å and four oxygen atoms of the two chelating sulfate anions. The vibrational spectra of K3VO2(SO4)2 are analyzed using these structural data.  相似文献   

16.
Solvent dynamics and polymer-solvent interactions in syndiotactic (s) polystyrene (PS)/ethylbenzene (PhEt) clathrates, as well as polymer-salt interactions in the poly(ethylene oxide) (PEO)/LiCF3SO3 complex, were characterized by solid state 1H and 13C NMR. 1H static and 1H MAS NMR spectra have shown that PhEt molecules in s-PS clathrates retain relatively large, but spatially anisotropic mobility. 13C CP/MAS (cross polarization/magic angle spinning) spectra and CP dynamics measured for s-PS-dg/PhEt system indicate that at least a part of PhEt molecules are intercalated between phenyl rings of s-PS. 13C CP/MAS NMR spectra show that PEO carbons in complex with LiCF3SO3 are more shielded in comparison to neat crystalline PEO. The results (distances) obtained from CP dynamics are in agreement with the published crystal structure of the PEO/LiCF3SO3 complex. 13C spin-lattice relaxation time measurements have shown that the mobility of PEO in the complex is lower than that in neat crystalline PEO.  相似文献   

17.
Preparation and Crystal Structure of Tetramethylammoniumbromide‐Bromine‐Sulfur Dioxide Adduct, [(CH3)4N]+Br�Br2�2SO2 Tetramethylammoniumtribromide forms with sulfur dioxide a salt which is characterized by vibrational spectroscopy and crystal structure analysis. [(CH3)4N]+Br�Br2�2SO2 crystallizes monoclinic in the space group P21/m with a = 657.4(5) pm, b = 2933.0(5) pm, c = 1462.2(5) pm, β = 91.241(5)° and four formula units in the unit cell. The crystal structure possesses bent infinite chains which consist of alternately arranged bromine and bromide ions. The bromide ions are connected to the molecules of bromine and sulfur dioxide by strong interactions forming a three dimensional network.  相似文献   

18.
Lü Jian 《中国化学》2011,29(2):283-287
The effect of metallic ions on the nitrolysis of DAPT [3,7‐diacetyl‐1,3,5,7‐tetraazabicyclo(3.3.1)nonane] and HA (hexamine) was investigated by experimental and theoretical approaches. The combinatorial reagent, M(NO?3)n/Ac2/NH4NO3 (M=Mg2+, Cu2+, Pb2+, Bi3+, Fe3+ and Zr4+), was found to be efficient in the experiment of the nitrolysis of DAPT. A key intermediate during the nitrolysis of DAPT was detected by 1H NMR. The formation mechanism of the intermediate was proposed and analyzed. Some discrepant results for the nitrolysis of DAPT and HA catalyzed by different metallic nitrates were explained based on hard‐soft and acid‐base principle and stabilized energy of ion‐complex. From the latter point of view, some cations with high polarizable ligands, e.g., OSO2CF3?, (CF3SO2)2N?, and (C4F9SO2)2N?, can increase the yields. Two newly designed catalysts, Cu[(CF3SO2)2N]2 and Cu[(C4F9SO2)2N]2, were tested to be highly efficient.  相似文献   

19.
La (III), Y(III) complexes wit11 diglycol aldehyde his-arginine (H2DAAR) and tetraglycol aldehyde bislysine (H2TALY) Schiff bases were synthesized. Thcy were characterized and formulated as La (H2DAAR) (NO3)3. 6H20 and Y (H2TALY) (NO3)3. 5H20 separately by elemental analyses. The obtained complexes were investigated in detail by high resolution solid state (HRSS)13C NMR using cross polarization, magic angle spinning (CPMAS). and total suppression of sidebands (TOSS) techniques. The results are supported by the liquid state 2D-’H-13C COSY NMR spectra. Some new information about the splitting peaks of13C for -CH = N- group and alkene carbon bands, etc. in HRSS13C NMR spectra are given.  相似文献   

20.
《中国化学会会志》2017,64(2):164-175
Two new silica‐supported vanadium(V) oxide nanocatalysts were synthesized from vanadyl(IV ) sulfate and vanadyl(IV ) oxalate by the sol–gel method. The nanocatalysts were characterized by Fourier transform infrared spectroscopy, thermogravimetry, X‐ray diffraction, X‐ray fluorescence, and field‐emission scanning electron microscopy. The average diameter of the nanocatalysts was ~16 nm. The catalytic activity of both nanocatalysts was investigated in the oxidation of sulfur dioxide (SO2 ) released from the roasting of copper sulfides to sulfur trioxide (SO3 ), and the results were compared with those obtained from a commercial BASF (Baden Aniline and Soda Factory) catalyst. It is worth mentioning that our nanocatalysts show higher activity than the commercial catalyst for the oxidation of SO2 to SO3 under similar reaction conditions. The suggested electrocatalytic mechanism for the oxidation of SO2 to SO3 was confirmed by computational studies. Our calculations indicate that the terminal VO bond of V2O5 is likely the most active site for the adsorption and oxidation of SO2 .  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