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.
Structures of stable compositions of sodium oxide cluster cations (NanOm+,n≤11) have been investigated by ion mobility mass spectrometry. Stoichiometric compositions series, Na(Na2O)(n-1)/2+ (n=3, 5, 7, 9, and 11), were observed as stable composition series, and NaO(Na2O)(n-1)/2+ series (n=5, 7, 9, and 11) were observed as secondary stable series in the mass spectra. To assign the structures of these cluster ion series, collision cross sections between the ions and helium buffer gas were determined experimentally from the ion mobility measurements. Theoretical collision cross sections were also calculated for optimized structures of these compositions. Finally, the structures of Na(Na2O)(n-1)/2+ and NaO(Na2O)(n-1)/2+ were assigned to those having similar structural frames for each n except for n=9. All bonds in the assigned structures of Na(Na2O)(n-1)/2+ were between sodium and oxygen. On the other hand, there was one O-O bond in addition to Na-O bonds in NaO(Na2O)(n-1)/2+. This result indicates that NaO(Na2O)(n-1)/2+ have a peroxide ion (O22-) as a substitute for an oxide ion (O2-) of Na(Na2O)(n-1)/2+. As a result, both stable series, Na(Na2O)(n-1)/2+ and NaO(Na2O)(n-1)/2+, are closed-shell compositions. These closed-shell characteristics have a strong influence on the stability of sodium oxide cluster cations.  相似文献   

2.
The title complex, {[Ni(C2H8N2)3][Na(NCS)3(H2O)]}n, con­sists of discrete [Ni(en)3]2+ dications (en is ethyl­enedi­amine) and polymeric [(H2O)0.5Na(NCS)3(H2O)0.5]n2n? anions. The compound crystallizes in space group Pc1. The NiII atom lies on a threefold axis and has a distorted octahedral coordination geometry. The Na+ cation also lies on a site with imposed crystallographic threefold symmetry and is coordinated by the thio­cyanate N atoms (the thio­cyanates are in general posi­tions), by one water mol­ecule with crystallographically imposed 32 symmetry and by a second water mol­ecule with crystallographically imposed symmetry. The unique Na atom thus has trigonal–bipyramidal coordination. The O atoms of the water mol­ecules bridge the Na+ cations to form one‐dimensional polymeric chains in the crystal structure. The [Ni(en)3]2+ dications are distributed around and between the chains and are linked to them via N—H?S hydrogen bonds.  相似文献   

3.
The energies of protonation and Na+ cationization of glycine (GLY) and its (GLY ? H + Na) salt in the gas phase were calculated using ab inltio calculations. The proton affinity of GLY, valued at the MP2/6–31G*//3-21G level, is 937 kJ mol?1. The amino function is confirmed to be the most favourable site of protonation: ‘proton affinities’ of the carbonyl and hydroxyl functions are calculated to be 75 and 180 kJ mol?1, respectively, lower than that of NH2 at the MP2/6-31G*//3–21G level. Calculations performed up to the MP2/6–31G*//3–21G level give the Na+ affinity of GLY as 189 kJ mol?1 and the H+ and Na+ affinities of (GLY – H + Na) as 1079 and 298 kJ mol?1, respectively. The geometries of all neutral and protonated species optimized with the 3–21G basis set are described. Both H* and Na+ cations complex preferably between the nitrogen atom and the carbonyl oxygen atom, leading to pseudo-five-membered ring structures in which Na? O and Na? N bonds lengths are greater than 2 Å.  相似文献   

4.
We studied the conditions for the photochemical formation of the NaZn excimer in the excited 22Π state using Na2(21Π u )+Zn→NaZn(22Π)+Na reaction. The Na-Zn vapor mixture was prepared in the heat-pipe oven with the well defined column density and temperature. The Na and Zn atom densities in the vapor mixture were controlled by the preparation of the alloy with different mole fraction ratios of the relevant components in the solid phase. The Na densities were determined from the total absorption coefficient at Na2 X-B and X-A bands. The cross section for photochemical formation of the NaZn in the 22Π state is estimated to be 17·10?16 cm2 for the laser excitation at 308 nm, measured relative to the cross section of 470·10?16 cm2 for collisional energy transfer Na2(21Π u )+Na→Na2(23Π g )+Na published by Mehdizadeh et al. [2].  相似文献   

5.
The electronic structure and redox properties of the highly oxidizing, isolable RuV?O complex [RuV(N4O)(O)]2+, its oxidation reactions with saturated alkanes (cyclohexane and methane) and inorganic substrates (hydrochloric acid and water), and its intermolecular coupling reaction have been examined by DFT calculations. The oxidation reactions with cyclohexane and methane proceed through hydrogen atom transfer in a transition state with a calculated free energy barrier of 10.8 and 23.8 kcal mol?1, respectively. The overall free energy activation barrier (ΔG=25.5 kcal mol?1) of oxidation of hydrochloric acid can be decomposed into two parts: the formation of [RuIII(N4O)(HOCl)]2+G=15.0 kcal mol?1) and the substitution of HOCl by a water molecule (ΔG=10.5 kcal mol?1). For water oxidation, nucleophilic attack on RuV?O by water, leading to O? O bond formation, has a free energy barrier of 24.0 kcal mol?1, the major component of which comes from the cleavage of the H? OH bond of water. Intermolecular self‐coupling of two molecules of [RuV(N4O)(O)]2+ leads to the [(N4O)RuIV? O2? RuIII(N4O)]4+ complex with a calculated free energy barrier of 12.0 kcal mol?1.  相似文献   

6.
Nonempirical quantum chemical method Hartree–Fock–Roothan LCAO SCF MO in a two-exponent Dunning basis with the use of an extended set of Gaussian functions by Huzinaga–Dunning with consideration of electron correlation according to the Meller–Plesset theory of excitations of the second order was used to study monohydrates of Li+, Na+, K+, and HCOO? ions. The indicated basis was supplemented with polarization functions of d-type on the O atom and of p-type on the hydrogen atom as well as with diffusion functions of p-type on the oxygen atom. It has been found that binding energies of the water molecule with Li+, Na+ appeared to be higher and with K+ lower than with HCOO? · H2O. Potential curve shapes of K+ + H2O and HCOO? + H2O reactions are shown to be similar. The molecular mechanism of K+ channel selectivity of an excitable membrane is explained on the basis of the obtained calculations.  相似文献   

7.
We treat the present work as an attempt to elucidate the mechanism of the oxidation reaction of the Cu atom by nitrous oxide based on our recent work (Kryachko, E. S.; Vinckier, C.; Nguyen, M. T. J Chem Phys 2001, 114, 7911) on the electron attachment to this molecule. We suggest that the title reaction in its Arrhenius regime occurs via the nonadiabatic electron transfer from Cu to the oxygen atom at the crossing of the potential energy surfaces Cu(4s 2S1/2) + N2O(X 1Σ+) and Cu+ + N2O?, where the latter is linked to the complex N2O? originated from the higher‐energy T‐shape N2O molecule and discovered in the aforementioned work. The calculations performed in the present work using a variety of quantum chemical methods support the proposed model. We also show the existence of other reaction pathways of the title reaction that, we believe, contribute to its non‐Arrhenius behavior observed experimentally at T > 1190 K. © 2002 Wiley Periodicals, Inc. Int J Quantum Chem, 2002  相似文献   

8.
The angular distribution of the chemiluminescent reaction Ba + N2O → BaO + N2 has been investigated by photographing directly the chemiluminescence from this reaction in a crossed beam experiment. It was found that the lifetime of the reactively scattered chemiluminescent BaO molecules is sufficiently long (≈ 10?s δ) to allow the observation of the angular distribution. From the dependence of this distribution on R and ? where R is the distance from the scattering center and ? the laboratory scattering angle, we conclude that under single collision conditions the chemiluminescence arises preferentially from highly excited vibrational-rotational levels of the A′1 Π state of BaO.  相似文献   

9.
Absolute gas phase Sn concentrations in the range 1 × 1013 ? [Sn] ? 1 × 1014 ml?1 have been determined utilizing a technique based on the rapid (at T ? 900 K) titration reaction Sn + NO2 → SnO + NO (k(900–1100 K) ≈ 1 × 10?10 ml molecule?1 s?1) and the chemiluminescent indicator reaction Sn + N2O → SnO + N2 + hv(SnO a3 Σ-X1Σ).  相似文献   

10.
In the title compound, [Na(C2H3N2O3)], the Na+ cation lies on a centre of inversion in space group P21/m and all the atoms of the anion lie on a mirror plane. Na is octahedrally coordinated by four O and two N atoms from six different anions and each anion is coordinated to six different Na+ cations, forming chains of confacial octahedra which link the anion layers. Within these layers, the individual anions are linked by both O—H?O and C—H?O hydrogen bonds.  相似文献   

11.
Herein, we report a theoretical and experimental study of the water‐gas shift (WGS) reaction on Ir1/FeOx single‐atom catalysts. Water dissociates to OH* on the Ir1 single atom and H* on the first‐neighbour O atom bonded with a Fe site. The adsorbed CO on Ir1 reacts with another adjacent O atom to produce CO2, yielding an oxygen vacancy (Ovac). Then, the formation of H2 becomes feasible due to migration of H from adsorbed OH* toward Ir1 and its subsequent reaction with another H*. The interaction of Ir1 and the second‐neighbouring Fe species demonstrates a new WGS pathway featured by electron transfer at the active site from Fe3+?O???Ir2+?Ovac to Fe2+?Ovac???Ir3+?O with the involvement of Ovac. The redox mechanism for WGS reaction through a dual metal active site (DMAS) is different from the conventional associative mechanism with the formation of formate or carboxyl intermediates. The proposed new reaction mechanism is corroborated by the experimental results with Ir1/FeOx for sequential production of CO2 and H2.  相似文献   

12.
We propose a new high temperature pathway for NO formation that involves the reaction of NNH with oxygen atoms. This reaction forms the HNNO* energized adduct via a rapid combination reaction; HNNO* then rapidly dissociates to NH + NO. The rate constant for O + NNH ? NH + NO is calculated via a QRRK chemical activation analysis to be 3.3 × 1014 T?0.23exp(+510/T) cm3 mol?1 s?1. This reaction sequence can be an important or even major route to NO formation under certain combustion conditions. The presence of significant quantities of NNH results from the reaction of H with N2. The H + N2 ? NNH reaction is only ca. 6 kcal/mol endothermic with a relatively low barrier. The reverse reaction, NNH dissociation, has been reported in the literature to be enhanced by tunneling. Our analysis of NNH dissociation indicates that tunneling dominates. We report a two-term rate constant for NNH dissociation: 3.0 × 108 + [M] {1.0 × 1013T0.5exp(?1540/T)} s?1. The first term accounts for pressure-independent tunneling from the ground vibrational state, while the second term accounts for collisional activation to higher vibration states from which tunneling can also occur. ([M] is the total concentration in units of mol cm?3.) Use of this dissociation rate constant and microscopic reversibility results in a large rate constant for the H + N2 reaction. As a result, we find that NNH ? H + N2 can be partially equilibrated under typical combustion conditions, resulting in NNH concentrations large enough for it to be important in bimolecular reactions. Our analysis of such reactions suggests that the reaction with oxygen atoms is especially important. © 1995 John Wiley & Sons, Inc.  相似文献   

13.
A laser flash photolysis–resonance fluorescence technique has been employed to investigate the kinetics of the reaction of ground state oxygen atoms, O(3PJ), with (CH3)2SO (dimethylsulfoxide) as a function of temperature (266–383 K) and pressure (20–100 Torr N2). The rate coefficient (kR1) for the O(3PJ) + (CH3)2SO reaction is found to be independent of pressure and to increase with decreasing temperature. The following Arrhenius expression adequately describes the observed temperature dependence: kR1(T) = (1.68 ± 0.76) × 10?12 exp[(445 ± 141)/T] cm3 molecule?1 s?1, where the uncertainties in Arrhenius parameters are 2σ and represent precision only. The absolute accuracy of each measured rate coefficient is estimated to be ±30%, and is limited predominantly by the uncertainties in measured (CH3)2SO concentrations. The observed temperature and pressure dependencies suggest that, as in the case of O(3PJ) reactions with CH3SH and (CH3)2S, reaction occurs by addition of O(3PJ) to the sulfur atom followed by rapid fragmentation of the energized adduct to products. The O(3PJ) + (CH3)2SO reaction is fast enough so that it could be a useful laboratory source of the CH3SO2 radical if this species is produced in significant yield. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 156–161, 2002; DOI 10.1002/kin.10040  相似文献   

14.
Heterogeneous recombination of O + CO → CO2 over a solid CO2 surface at 77 K was investigated. A modified discharge flow setup was used to generate low O atom concentrations by the reaction N + NO → N2 + O(3P). The O atom concentrations were measured upstream and downstream of the solid CO2 substrate using resonance fluorescence by monitoring the unresolved 130.3 nm triplet transition 3S1 ? 3P2,1,0 at the two fixed points. CO2 formed was determined by measuring the β activity from C14O2 produced from CO containing C14O as a reactant gas. The CO2 formation was found to be first order in CO and independent of O atom concentration over the entire range of 4.3 × 1012 to 1.9 × 1014 cm?3 and 1.2 × 1011 to 5.6 × 1012 cm?3 for CO and O respectively. The first order recombination coefficient, λCO was found to be 1.4 (±.38) × 10?5.  相似文献   

15.
Three complexes, Na4[DyIII(dtpa)(H2O)]2?·?16H2O, Na[DyIII(edta)(H2O)3]?·?3.25H2O and Na3[DyIII (nta)2(H2O)]?·?5.5H2O, have been synthesized in aqueous solution and characterized by FT–IR, elemental analyses, TG–DTA and single-crystal X-ray diffraction. Na4[DyIII(dtpa)(H2O)]2?·?16H2O crystallizes in the monoclinic system with P21/n space group, a?=?18.158(10)?Å, b?=?14.968(9)?Å, c?=?20.769(12)?Å, β?=?108.552(9)°, V?=?5351(5)?Å3, Z?=?4, M?=?1517.87?g?mol?1, D c?=?1.879?g?cm?3, μ?=?2.914?mm?1, F(000)?=?3032, and its structure is refined to R 1(F)?=?0.0500 for 9384 observed reflections [I?>?2σ(I)]. Na[DyIII(edta)(H2O)3]?·?3.25H2O crystallizes in the orthorhombic system with Fdd2 space group, a?=?19.338(7)?Å, b?=?35.378(13)?Å, c?=?12.137(5)?Å, β?=?90°, V?=?8303(5)?Å3, Z?=?16, M?=?586.31?g?mol?1, D c?=?1.876?g?cm?3, μ?=?3.690?mm?1, F(000)?=?4632, and its structure is refined to R 1(F)?=?0.0307 for 4027 observed reflections [I?>?2σ(I)]. Na3[DyIII(nta)2(H2O)]?·?5.5H2O crystallizes in the orthorhombic system with Pccn space group, a?=?15.964(12)?Å, b?=?19.665(15)?Å, c?=?14.552(11)?Å, β?=?90°, V?=?4568(6)?Å3, Z?=?8, M?=?724.81?g?mol?1, D c?=?2.102?g?cm?3, μ?=?3.422?mm?1, F(000)?=?2848, and its structure is refined to R 1(F)?=?0.0449 for 4033 observed reflections [I?>?2?σ(I)]. The coordination polyhedra are tricapped trigonal prism for Na4[DyIII(dtpa)(H2O)]2?·?16H2O and Na3[DyIII(nta)2(H2O)]?·?5.5H2O, but monocapped square antiprism for Na[DyIII(edta)(H2O)3]?·?3.25H2O. The crystal structures of these three complexes are completely different from one another. The three-dimensional geometries of three polymers are 3-D layer-shaped structure for Na4[DyIII(dtpa)(H2O)]2?·?16H2O, 1-D zigzag type structure for Na[DyIII(edta)(H2O)3]?·?3.25H2O and a 2-D parallelogram for Na3[DyIII(nta)2(H2O)]?·?5.5H2O. According to thermal analyses, the collapsing temperatures are 356°C for Na4[DyIII(dtpa)(H2O)]2?·?16H2O, 371°C for Na[DyIII(edta)(H2O)3]?·?3.25H2O and 387°C for Na3[DyIII(nta)2(H2O)]?·?5.5H2O, which indicates that their crystal structures are very stable.  相似文献   

16.
The effects of charged species on proton‐coupled electron‐transfer (PCET) reaction should be of significance for understanding/application of important chemical and biological PCET systems. Such species can be found in proximity of activated complex in a PCET reaction, although they are not involved in the charge transfer process. Reported here is the first study of the above‐mentioned effects. Here, the effects of Na+, K+, Li+, Ca2+, Mg2+, and Me4N+ observed in PCET reaction of ascorbate monoanions with hexacyanoferrate(III) ions in H2O reveal that, in presence of ions, this over‐the‐barrier reaction entered into tunneling regime. The observations are: a) dependence of the rate constant on the cation concentration, where the rate constant is 71 (at I = 0.0023), and 821 (at 0.5M K+), 847 (at 1.0M Na+), and 438 M ?1 s?1 (at 0.011M Ca2+); b) changes of kinetic isotope effect (KIE) in the presence of ions, where kH/kD=4.6 (at I = 0.0023), and 3.4 (in the presence of 0.5M K+), 3.3 (at 1.0M Na+), 3.9 (at 0.001M Ca2+), and 3.9 (at 0.001M Mg2+), respectively; c) the isotope effects on Arrhenius pre‐factor where AH/AD=0.97 (0.15) in absence of ions, and 2.29 (0.60) (at 0.5M Na+), 1.77 (0.29) (at 1.0M Na+), 1.61 (0.25) (at 0.5M K+), 0.42 (0.16) (at 0.001M Ca2+) and 0.16 (0.19) (at 0.001M Mg2+); d) isotope differences in the enthalpies of activation in H2O and in D2O, where ΔΔH?(D,H)=3.9 (0.4) kJ mol?1 in the absence of cations, 1.3 (0.6) at 0.5M Na+, 1.8 (0.4) at 0.5M K+, 1.5 (0.4) at 1.0M Na+, 5.5 (0.9) (at 0.001M Ca2+), and 7.9 (2.8) (at 0.001M Mg2+) kJ mol?1; e) nonlinear proton inventory in reaction. In the H2O/dioxane 1 : 1, the observed KIE is 7.8 and 4.4 in the absence and in the presence of 0.1M K+, respectively, and AH/AD=0.14 (0.03). The changes when cations are present in the reaction are explained in terms of termolecular encounter complex consisting of redox partners, and the cation where the cation can be found in a near proximity of the reaction‐activated complex thus influencing the proton/electron double tunneling event in the PCET process. A molecule of H2O is involved in the transition state. The resulting ‘configuration’ is more ‘rigid’ and more appropriate for efficient tunneling with Na+ or K+ (extensive tunneling observed), i.e., there is more precise organized H transfer coordinate than in the case of Ca2+ and Mg2+ (moderate tunneling observed) in the reaction.  相似文献   

17.
Atomic resonance absorption spectroscopy has been used to investigate the thermal decomposition of N2O by monitoring the formation of O atoms behind reflected shock waves in the temperature range 1490–2490 K and at total pressures from 58 to 347 kPa, by using the mixtures of N2O highly diluted in Ar. For the chosen experimental conditions, the rate coefficient k1,0 for the reaction N2O + Ar → N2 + O + Ar had the greatest effect on the O atom concentration increase, so this reaction rate constant could be deduced by comparison between experiment and computed simulation. In the actual temperature range, we found k1,0 (cm3 mol?1s?1) = 7.2 × 1014 exp(?28878/T(K)), with an overall uncertainty evaluated to be less than 20%, by considering all the parameters, which contributed to uncertainties in the rate constant determination. The possible absorption at the O triplet emission line of N2O has been investigated. The absorption cross section of N2O at the O line has been estimated and taken into account for the determination of k1,0 at high concentrations of N2O and at temperatures lower than 1850 K. The effect of the presence of impurities like H2O on rate constant determination has been examined and was found to be negligible. The choice of the rate coefficient for the consumption of O atoms by reaction with N2O and that of the high‐pressure limiting rate coefficients k1,∞ were also discussed. The rate constant reported in the present study was compared with the literature values and was found to be overall higher than those determined experimentally by other teams in the last decade. Finally, the effect of the modified constant value on reaction rate of diluted Ar–N2O mixtures and H2–N2O–Ar systems was investigated. In the temperature range 1500–2500 K, the use of the rate constant deduced from this study has led to a better prediction of N2O decomposition and N2O reduction by H2 than with lower rate constants proposed in the literature. © 2009 Wiley Periodicals, Inc. Int J Chem Kinet 41: 357–375, 2009  相似文献   

18.
Smog chamber/FTIR techniques were used to study the Cl atom initiated oxidation of CH2FOCH2F in 700 Torr of N2/O2 at 296 K. Relative rate techniques were used to measure k(Cl + CH2FOCH2F) = (4.6 ± 0.7) × 10?13 and k(Cl + CH2FOC(O)F) = (2.9 ± 0.8) × 10?15 (in units of cm3 molecule?1 s?1). Three competing fates for alkoxy radical CH2FOCHFO· formed in the self‐reaction of the corresponding peroxy radicals were identified. In 1 atm of air at 296 K, 48 ± 3% of CH2FOCHFO· radicals decompose via C? O bond scission, 21 ± 4% react with O2, and 31 ± 4% undergo hydrogen atom elimination. Chemical activation effects were observed for CH2FOCHFO· radicals formed in the CH2FOCHFOO· + NO reaction. Infrared spectra of CH2FOC(O)F and FC(O)OC(O)F, which are produced during the Cl atom initiated oxidation of CH2FOCH2F, are presented. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 139–147, 2002; DOI 10.1002/kin.10038  相似文献   

19.
The title compound, tricaesium sodium iron(III) μ3‐oxido‐hexa‐μ2‐sulfato‐tris[aquairon(III)] pentahydrate, Cs2.91Na1.34Fe3+0.25[Fe3O(SO4)6(H2O)3]·5H2O, belongs to the family of Maus's salts, K5[Fe3O(SO4)6(H2O)3]·6H2O, which is based on the triaqua‐μ3‐oxido‐hexa‐μ‐sulfato‐triferrate(III) anion, [Fe3O(SO4)6(H2O)3]5−, with Fe in a characteristically distorted octahedral coordination environment, sharing a common corner via an oxide O atom. Cs in four different cation sites, Na in three different cation sites and five water molecules link the anions in three dimensions and set up a crystal structure in which those parts parallel to (001) and within 0.05 < z < 0.95 have a distinct trigonal pseudosymmetry, whereas the cation arrangement and bonding near z∼ 0 generate a clear‐cut noncentrosymmetric polar edifice with the monoclinic space group C2. The structure shows some cation disorder in the region near z ∼ , where one Na atom in octahedral coordination is partly substituted by Fe3+, and a Cs atom is substituted by small amounts of Na on a separate nearby site. One Na atom, located on a twofold axis at z = 0 and tetrahedrally coordinated by four sulfate O atoms of two [Fe3O(SO4)6(H2O)3]5− units, plays a key role in generating the noncentrosymmetric structure. Three of the seven different cation sites are on twofold axes (one Na+ site and two Cs+ sites), and all other atoms of the structure are in general positions.  相似文献   

20.
The reactions of ground-state oxygen atoms with carbonothioicdichloride, carbonothioicdifluoride, and tetrafluoro-1,3-dithietane have been studied in a crossed molecular jet reactor in order to determine the initial reaction products and in a fast-flow reactor in order to determine their overall rate constants at temperatures between 250 and 500 K. These rate constants are??(O + C2CS) =(3.09 ± 0.54) × 10?11 exp(+115 ± 106 cal/mol/RT),??(O + F2CS) = (1.22 ± 0.19) × 10?11 exp(-747 ± 95 cal/mol/RT), and??(O + F4C2S2) = (2.36 ± 0.52) × 10?11 exp(-1700 ± 128 cal/mol/RT) cm3/molec˙sec. The detected reaction products and their rate constants indicate that the primary reaction mechanism is the electrophilic addition of the oxygen atom to the sulfur atom contained in the reactant molecule to form an energy-rich adduct which then decomposes by C-S bond cleavage.  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