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.
Five [BW12O40]5 ? -containing metal-organic frameworks (MOFs), {K[Ln(H2O)4(pdc)]4}[BW12O40]·2H2O (Ln=La 1 and Ce 2, H2pdc = pyridine-2,6-dicarboxylate) and {K[Ln(H2O)3(pdc)]4}[BW12O40]·6H2O (Ln=Tb 3, Dy 4 and Er 5), are synthesized hydrothermally. Ln3+ and pdc2? are built into 3-D MOF segments containing large channels and cavities which are occupied by [BW12O40]5 ? . Herein, we report MOF-containing polyoxometalates (POMs) as photocatalysts to oxidize thiophene with O2 as the oxidant. Photo-excited state species ([BW12O40]5??) are generated under UV irradiation and then H2O is oxidized into OH˙ radicals and O2 is reduced into O2 2?. The active oxygen species (O2 2?, OH˙) oxidize thiophene in the presence of photocatalysts and SO3, CO2, and H2O are obtained as photoproducts. ESR measurements provide evidence that OH˙ species are generated during photocatalysis.  相似文献   

2.
The opto-acoustic spectrum of I2 in the presence of various quenching gases — NO, O2, CH3I, SO2, C3HS, N2, and He — has been studied. Of these, the I2/O2 spectrum is quite different due to the near-resonant energy transfer I(2P12) + O2(3Σ) → I(2P32) + O2(IΔ), wherein the resistance of the O2((IΔ) species to collisional relaxation severely distorts the acoustic signal. The photochemical production of excited 2P12 iodine atoms commences at wavelengths considerably longer than the dissociation limit of the I2B? state.  相似文献   

3.
The Li, Rb and Cs complexes with the herbicide (2,4‐dichlorophenoxy)acetic acid (2,4‐D), namely poly[[aqua[μ3‐(2,4‐dichlorophenoxy)acetato‐κ3O1:O1:O1′]lithium(I)] dihydrate], {[Li(C8H5Cl2O3)(H2O)]·2H2O}n, (I), poly[μ‐aqua‐bis[μ3‐(2,4‐dichlorophenoxy)acetato‐κ4O1:O1′:O1′,Cl2]dirubidium(I)], [Rb2(C8H5Cl2O3)2(H2O)]n, (II), and poly[μ‐aqua‐bis[μ3‐(2,4‐dichlorophenoxy)acetato‐κ5O1:O1′:O1′,O2,Cl2]dicaesium(I)], [Cs2(C8H5Cl2O3)2(H2O)]n, (III), respectively, have been determined and their two‐dimensional polymeric structures are described. In (I), the slightly distorted tetrahedral LiO4 coordination involves three carboxylate O‐atom donors, of which two are bridging, and a monodentate aqua ligand, together with two water molecules of solvation. Conjoined six‐membered ring systems generate a one‐dimensional coordination polymeric chain which extends along b and interspecies water O—H...O hydrogen‐bonding interactions give the overall two‐dimensional layers which lie parallel to (001). In hemihydrate complex (II), the irregular octahedral RbO5Cl coordination about Rb+ comprises a single bridging water molecule which lies on a twofold rotation axis, a bidentate Ocarboxy,Cl‐chelate interaction and three bridging carboxylate O‐atom bonding interactions from the 2,4‐D ligand. A two‐dimensional coordination polymeric layer structure lying parallel to (100) is formed through a number of conjoined cyclic bridges, including a centrosymmetric four‐membered Rb2O2 ring system with an Rb...Rb separation of 4.3312 (5) Å. The coordinated water molecule forms intralayer aqua–carboxylate O—H...O hydrogen bonds. Complex (III) comprises two crystallographically independent (Z′ = 2) irregular CsO6Cl coordination centres, each comprising two O‐atom donors (carboxylate and phenoxy) and a ring‐substituted Cl‐atom donor from the 2,4‐D ligand species in a tridentate chelate mode, two O‐atom donors from bridging carboxylate groups and one from a bridging water molecule. However, the two 2,4‐D ligands are conformationally very dissimilar, with one phenoxyacetate side chain being synclinal and the other being antiperiplanar. The minimum Cs...Cs separation is 4.4463 (5) Å. Structure extension gives coordination polymeric layers which lie parallel to (001) and are stabilized by intralayer water–carboxylate O—H...O hydrogen bonds.  相似文献   

4.
Pulse radiolysis techniques were used to measure the gas phase UV absorption spectra of the title peroxy radicals over the range 215–340 nm. By scaling to σ(CH3O2)240 nm = (4.24 ± 0.27) × 10?18, the following absorption cross sections were determined: σ(HO2)240 nm = 1.29 ± 0.16, σ(C2H5O2)240 nm = 4.71 ± 0.45, σ(CH3C(O)CH2O2)240 nm = 2.03 ± 0.22, σ(CH3C(O)CH2O2)230 nm = 2.94 ± 0.29, and σ(CH3C(O)CH2O2)310 nm = 1.31 ± 0.15 (base e, units of 10?18 cm2 molecule?1). To support the UV measurements, FTIR‐smog chamber techniques were employed to investigate the reaction of F and Cl atoms with acetone. The F atom reaction proceeds via two channels: the major channel (92% ± 3%) gives CH3C(O)CH2 radicals and HF, while the minor channel (8% ± 1%) gives CH3 radicals and CH3C(O)F. The majority (>97%) of the Cl atom reaction proceeds via H atom abstraction to give CH3C(O)CH2 radicals. The results are discussed with respect to the literature data concerning the UV absorption spectra of CH3C(O)CH2O2 and other peroxy radicals. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 283–291, 2002  相似文献   

5.
The environment of H3O+, H2O, HF and F species (non-bonded to metals) is considered in fluoride metalates which crystallise from the (Al(OH)3, Cr(OH)3, FeF3, ZrF4, Ta2O5)-tren-HFaq·-ethanol systems (microwave heating at 190 °C during 1 h). The presence of (H3O)(H2O)6+ clusters or H3O+ cations, of isolated or associated H2O molecules, of (HF2) and F anions is evidenced. The thermal stability of the solids depends strongly on the nature of the hydrogen-bonded species associated with the preceding cations or anions and on the formation of water ribbons or layers.  相似文献   

6.
The substituent methoxy group at the phenyl ortho position in the title compound, C27H22O3, has an insignificant effect on the length of the Csp3—O bond and on the non‐planarity of the pyran ring. The cause of the changes in the photochemical properties is discussed.  相似文献   

7.
The kinetics of the deactivation of O2(1Σg+) is studied in real time. O2(1Σg+) is generated in this system by the O(1D) + O2 reaction following O3laser flash photolysis in the presence of excess O2, and it is monitored by its characteristic emission band at 762 nm. Quenching rate constants were obtained for O2, O3, N2, CO2, H2O, CF4and the rare gases. Since O(1D) is the precursor for the formation of O2(1Σg+), the addition of an O(1D) quencher effectively lowers the initial concentration of O2(1Σg+). By measuring the initial intensity of the 762 nm fluorescence signal, the relative quenching efficiencies were determined for O(1D) quenching by N2, CO2, Xe, and Kr with respect to O2; the results are in good agreement with literature values.  相似文献   

8.
New obtained is K2Cd2O3 (brownish red), which – according to single crystal work [487 h 01–h 41, Mo? Kα, R = 9.73%, R′ = 10.76%] – crystallises monoclinic with a = 6.417, b = 6.723, c = 6.586 Å, β = 116.0° in P21/c–C [K+, Cd2+ and O(1)2? in 4(e), O(2)2? in 2(a)] and is isotypic with Na2Zn2O3. [Parameters see text]. The Madelung Part of Lattice Energy is calculated and discussed.  相似文献   

9.
A simple and effective synthetic route to homo‐ and heteroleptic rare‐earth (Ln = Y, La and Nd) complexes with a tridentate Schiff base anion has been demonstrated using exchange reactions of rare‐earth chlorides with in‐situ‐generated sodium (E)‐2‐{[(2‐methoxyphenyl)imino]methyl}phenoxide in different molar ratios in absolute methanol. Five crystal structures have been determined and studied, namely tris(2‐{[(2‐methoxyphenyl)imino]methyl}phenolato‐κ3O1,N,O2)lanthanum, [La(C14H12NO2)3], ( 1 ), tris(2‐{[(2‐methoxyphenyl)imino]methyl}phenolato‐κ3O1,N,O2)neodymium tetrahydrofuran disolvate, [La(C14H12NO2)3]·2C4H8O, ( 2 )·2THF, tris(2‐{[(2‐methoxyphenyl)imino]methyl}phenolato)‐κ3O1,N,O23O1,N,O22N,O1‐yttrium, [Y(C14H12NO2)3], ( 3 ), dichlorido‐1κCl,2κCl‐μ‐methanolato‐1:2κ2O:O‐methanol‐2κO‐(μ‐2‐{[(2‐methoxyphenyl)imino]methyl}phenolato‐1κ3O1,N,O2:2κO1)bis(2‐{[(2‐methoxyphenyl)imino]methyl}phenolato)‐1κ3O1,N,O2;2κ3O1,N,O2‐diyttrium–tetrahydrofuran–methanol (1/1/1), [Y2(C14H12NO2)3(CH3O)Cl2(CH4O)]·CH4O·C4H8O, ( 4 )·MeOH·THF, and bis(μ‐2‐{[(2‐methoxyphenyl)imino]methyl}phenolato‐1κ3O1,N,O2:2κO1)bis(2‐{[(2‐methoxyphenyl)imino]methyl}phenolato‐2κ3O1,N,O2)sodiumyttrium chloroform disolvate, [NaY(C14H12NO2)4]·2CHCl3, ( 5 )·2CHCl3. Structural peculiarities of homoleptic tris(iminophenoxide)s ( 1 )–( 3 ), binuclear tris(iminophenoxide) ( 4 ) and homoleptic ate tetrakis(iminophenoxide) ( 5 ) are discussed. The nonflat Schiff base ligand displays μ2‐κ3O1,N,O2O1 bridging, and κ3O1,N,O2 and κ2N,O1 terminal coordination modes, depending on steric congestion, which in turn depends on the ionic radii of the rare‐earth metals and the number of coordinated ligands. It has been demonstrated that interligand dihedral angles of the phenoxide ligand are convenient for comparing steric hindrance in complexes. ( 4 )·MeOH has a flat Y2O2 rhomboid core and exhibits both inter‐ and intramolecular MeO—H…Cl hydrogen bonding. Catalytic systems based on complexes ( 1 )–( 3 ) and ( 5 ) have demonstrated medium catalytic performance in acrylonitrile polymerization, providing polyacrylonitrile samples with narrow polydispersity.  相似文献   

10.
The fluoropolytungstates [H2W12F2O38]4? and [HW12F3O37]4?, which are of the metatungstate type with the fluoride ions occupying inner sites, lose fluoride ion and form more higly charged species [H2W12FO39]5? and [HW12F2O38]5? in aqueous solution about pH 3 in media of suitable ionic strength. Kinetic results are presented here consistent with a mechanism involving less condensed species containing 11 tungsten and 1, 2 or 3 fluoride atoms. Involvement of such species is supported by the isolation of the aluminotungstates [HW11AlF3O36(H2O)] (TMA)52, 30H2O and [H2W11AlF2O37(H2O)](TMA)5, 17H2O.  相似文献   

11.
Iron(II), (Fe(H2O)62+, (FeII) participates in many reactions of natural and biological importance. It is critically important to understand the rates and the mechanism of FeII oxidation by dissolved molecular oxygen, O2, under environmental conditions containing bicarbonate (HCO3), which exists up to millimolar concentrations. In the absence and presence of HCO3, the formation of reactive oxygen species (O2, H2O2, and HO⋅) in FeII oxidation by O2 has been suggested. In contrast, our study demonstrates for the first time the rapid generation of carbonate radical anions (CO3) in the oxidation of FeII by O2 in the presence of bicarbonate, HCO3. The rate of the formation of CO3 may be expressed as d[CO3]/dt=[FeII[[O2][HCO3]2. The formation of reactive species was investigated using 1H nuclear magnetic resonance (1H NMR) and gas chromatographic techniques. The study presented herein provides new insights into the reaction mechanism of FeII oxidation by O2 in the presence of bicarbonate and highlights the importance of considering the formation of CO3 in the geochemical cycling of iron and carbon.  相似文献   

12.
The new tetranuclear alkoxide hexa‐μ2‐isopropoxy‐1:2κ4O;1:3κ4O;1:4κ4O‐hexaisopropoxy‐2κ2O,3κ2O,4κ2O‐trialumin­ium(III)­neodymium(III), [Nd{Al(C3H7O)4}3], has a metal–oxy­gen NdAl3O12 core which consists of four metal atoms arranged in an approximately planar triangular geometry. The central Nd atom is six‐coordinated by O atoms and the Al atoms are four‐coordinated by O atoms.  相似文献   

13.
The thermal decomposition of trifluoromethoxycarbonyl peroxy nitrate, CF3OC(O)O2NO2, has been studied between 278 and 306 K at 270 mbar total pressure using He as a diluent gas. The pressure dependence of the reaction was also studied at 292 K between 1.2 and 270 mbar total pressure. The rate constant reaches its high‐pressure limit at 70 mbar. The first step of the decomposition leads to CF3OC(O)O2 and NO2 formation, that is, CF3OC(O)O2NO2 + M ? CF3OC(O)O2 + NO2 + M (k1, k?1). Reaction (?1) was prevented by adding an excess of NO that reacts with the peroxy radical intermediate and leads to carbonyl fluoride (CF2O), carbon dioxide (CO2), nitrogen dioxide (NO2), and small quantities of CF3OC(O)O2C(O)OCF3. The kinetics of reaction (1) was determined by following the loss of CF3OC(O)O2NO2 via IR spectroscopy. The temperature dependence of the decomposition follows the equation k1(T) = 1.0 × 1016 e?((111±3)/(RT)) for the exponential term expressed in kJ mol?1. The values obtained for the kinetic parameters such as k1 at 298 K, the activation energy (Ea), and the preexponential factor (A) are compared with literature data for other acyl peroxy nitrates. The atmospheric thermal stability of CF3OC(O)O2NO2 and its dependence with altitude is discussed. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 40: 831–838, 2008  相似文献   

14.
A one‐pot template condensation of 2‐(2‐(dicyanomethylene)hydrazinyl)benzenesulfonic acid (H2L1, 1 ) or 2‐(2‐(dicyanomethylene)hydrazinyl)benzoic acid (H2L2, 2 ) with methanol (a), ethylenediamine (b), ethanol (c) or water (d) on copper(II), led to a variety of metal complexes, that is, mononuclear [Cu(H2O)2O1N2 L1a] ( 3 ) and [Cu(H2O)(κO1N3 L1b)] ( 4 ), tetranuclear [Cu4(1 κO1N2:2 κO1 L2a)3‐(1 κO1, κN2:2 κO2 L2a)] ( 5 ), [Cu2(H2O)(1 κO1, κN2:2 κO1 L2c)‐(1 κO1,1 κN2:2 κO1,2 κN1‐ L2c)]2 ( 6 ) and [Cu2(H2O)2O1N2‐ L1dd)‐(1 κO1N2:2 κO1 L1dd)(μ‐H2O)]2 ? 2 H2O ( 7? 2 H2O), as well as polymer‐ ic [Cu(H2O)(κO1,1 κN2:2 κN1 L1c)]n ( 8 ) and [Cu(NH2C2H5)(κO1,1 κN2:2 κN1L2a)]n ( 9 ). The ligands 2‐SO3H‐C6H4‐(NH)N?C{(CN)[C(NH2)‐(?NCH2CH2NH2)]} (H2L1b, 10 ), 2‐CO2H‐C6H4‐(NH)N?{C(CN)[C(OCH3)‐(?NH)]} (H2L2a, 11 ) and 2‐SO3H‐C6H4‐(NH)N?C{C(?O)‐(NH2)}2 (H2L1dd, 12 ) were easily liberated upon respective treatment of 4 , 5 and 7 with HCl, whereas the formation of cyclic zwitterionic amidine 2‐(SO3?)? C6H4? N?NC(? C?(NH+)CH2CH2NH)(?CNHCH2CH2NH) ( 13 ) was observed when 1 was treated with ethylenediamine. The hydrogen bond‐induced E/Z isomerization of the (HL1d)? ligand occurs upon conversion of [{Na(H2O)2(μ‐H2O)2}(HL1d)]n ( 14 ) to [Cu(H2O)6][HL1d]2 ? 2 H2O ( 15 ) and [{CuNa(H2O)‐(κN1,1 κO2:2 κO1 L1d)2}K0.5(μ‐O)2]n ? H2O ( 16 ). The synthesized complexes 3 – 9 are catalyst precursors for both the selective oxidation of primary and secondary alcohols (to the corresponding carbonyl compounds) and the following diastereoselective nitroaldol (Henry) reaction, with typical yields of 80–99 %.  相似文献   

15.
The crystal structures of five new transition‐metal complexes synthesized using thiazole‐2‐carboxylic acid (2‐Htza), imidazole‐2‐carboxylic acid (2‐H2ima) or 1,3‐oxazole‐4‐carboxylic acid (4‐Hoxa), namely diaquabis(thiazole‐2‐carboxylato‐κ2N,O)cobalt(II), [Co(C4H2NO2S)2(H2O)2], 1 , diaquabis(thiazole‐2‐carboxylato‐κ2N,O)nickel(II), [Ni(C4H2NO2S)2(H2O)2], 2 , diaquabis(thiazole‐2‐carboxylato‐κ2N,O)cadmium(II), [Cd(C4H2NO2S)2(H2O)2], 3 , diaquabis(1H‐imidazole‐2‐carboxylato‐κ2N3,O)cobalt(II), [Co(C4H2N2O2)2(H2O)2], 4 , and diaquabis(1,3‐oxazole‐4‐carboxylato‐κ2N,O4)cobalt(II), [Co(C4H2NO3)2(H2O)2], 5 , are reported. The influence of the nature of the heteroatom and the position of the carboxyl group in relation to the heteroatom on the self‐assembly process are discussed based upon Hirshfeld surface analysis and used to explain the observed differences in the single‐crystal structures and the supramolecular frameworks and topologies of complexes 1 – 5 .  相似文献   

16.
The structures of the anhydrous 1:1 proton‐transfer compounds of 4,5‐dichlorophthalic acid (DCPA) with the monocyclic heteroaromatic Lewis bases 2‐aminopyrimidine, 3‐(aminocarbonyl)pyridine (nicotinamide) and 4‐(aminocarbonyl)pyridine (isonicotinamide), namely 2‐aminopyrimidinium 2‐carboxy‐4,5‐dichlorobenzoate, C4H6N3+·C8H3Cl2O4, (I), 3‐(aminocarbonyl)pyridinium 2‐carboxy‐4,5‐dichlorobenzoate, C6H7N2O+·C8H3Cl2O4, (II), and the unusual salt adduct 4‐(aminocarbonyl)pyridinium 2‐carboxy‐4,5‐dichlorobenzoate–methyl 2‐carboxy‐4,5‐dichlorobenzoate (1/1), C6H7N2O+·C8H3Cl2O4·C9H6Cl2O4, (III), have been determined at 130 K. Compound (I) forms discrete centrosymmetric hydrogen‐bonded cyclic bis(cation–anion) units having both R22(8) and R12(4) N—H...O interactions. In (II), the primary N—H...O‐linked cation–anion units are extended into a two‐dimensional sheet structure via amide–carboxyl and amide–carbonyl N—H...O interactions. The structure of (III) reveals the presence of an unusual and unexpected self‐synthesized methyl monoester of the acid as an adduct molecule, giving one‐dimensional hydrogen‐bonded chains. In all three structures, the hydrogen phthalate anions are essentially planar with short intramolecular carboxyl–carboxylate O—H...O hydrogen bonds [O...O = 2.393 (8)–2.410 (2) Å]. This work provides examples of low‐dimensional 1:1 hydrogen‐bonded DCPA structure types, and includes the first example of a discrete cyclic `heterotetramer.' This low dimensionality in the structures of the 1:1 aromatic Lewis base salts of the parent acid is generally associated with the planar DCPA anion species.  相似文献   

17.
The structures of three copper‐containing complexes, namely (benzoato‐κ2O,O′)[(E)‐2‐({[2‐(diethylamino)ethyl]imino}methyl)phenolato‐κ3N,N′,O]copper(II) dihydrate, [Cu(C7H5O2)(C13H19N2O)]·2H2O, 1 , [(E)‐2‐({[2‐(diethylamino)ethyl]imino}methyl)phenolato‐κ3N,N′,O](2‐phenylacetato‐κ2O,O′)copper(II), [Cu(C8H7O2)(C13H19N2O)], 2 , and bis[μ‐(E)‐2‐({[3‐(diethylamino)propyl]imino}methyl)phenolato]‐κ4N,N′,O:O4O:N,N′,O‐(μ‐2‐methylbenzoato‐κ2O:O′)copper(II) perchlorate, [Cu2(C8H7O2)(C12H17N2O)2]ClO4, 3 , have been reported and all have been tested for their activity in the oxidation of d ‐galactose. The results suggest that, unlike the enzyme galactose oxidase, due to the precipitation of Cu2O, this reaction is not catalytic as would have been expected. The structures of 1 and 2 are monomeric, while 3 consists of a dimeric cation and a perchlorate anion [which is disordered over two orientations, with occupancies of 0.64 (4) and 0.36 (4)]. In all three structures, the central Cu atom is five‐coordinated in a distorted square‐pyramidal arrangment (τ parameter of 0.0932 for 1 , 0.0888 for 2 , and 0.142 and 0.248 for the two Cu centers in 3 ). In each species, the environment about the Cu atom is such that the vacant sixth position is open, with very little steric crowding.  相似文献   

18.
Abstract

The reactions of Mo(CO)6 and W(CO)6 with HCl(g) in the presence of 12-crown-4 and H2O have been investigated in toluene. For both reactions, two products were isolated, depending on the oxidation of the metal center. For molybdenum, the MoIII species, [H3O+ · 12-crown-4]3[Mo2Cl9 3-], 1, was obtained from the liquid clathrate layer in the reaction mixture. Upon air oxidation of the reaction mixture, the Mov complex, [H7O3 ? · H4O2 + · (12-crown-4)2][MoOCl4(H2O)?]2, 2, rapidly formed. For tungsten, the WII species, [(H5O2 +)2 · 12-crown-4][W(CO)4Cl? 3]2, 3, deposited from the liquid clathrate layer which upon oxidation formed the Wv complex, [H3O+· 12-crown-4][WOCl4(H2O)?], 4. These reactions were promoted by UV radiation and formed liquid clathrates almost immediately upon reaction. X-ray crystal structures were performed on each compound. Complexes 1 and 4 have H3O+ oxonium ions involved in complex hydrogen bonded arrays with the 12-crown-4 acceptor molecules. The H5O2 + oxonium ions in 2 and 3 contain extremely short O…O separations, equivalent to the shortest O-H…O bonds known. Also isolated in complex 2 was the H7O3 + oxonium ion which contains an unusual linear O…O…O core.  相似文献   

19.
Three new extended iron‐containing heteropolytungstates were synthesized and structurally characterized: K6[{FeII(H2O)4}2(H2W12O42)]·15H2O ( 1 ), Na5[{Fe(H2O)3}2{Fe(H2O)4}0.5(H2W12O42)]·30H2O ( 2 ) and (H3O)+2[{Fe(H2O)4Fe(H2O)3}2(H2W12O42)]·20H2O ( 3 ). 1 and 3 crystallize in the monoclinic system, space group P21/n with a = 14.9967(5), b = 10.3872(3), c = 18.8237(6)Å, β = 93.407(1)°, V = 2927.1(2)Å3 and Dc = 4.151 g cm—3 for 1 , and space group P21/c with a = 12.1794(4), b = 22.4938(4), c = 11.6941(3) Å, β = 105.731(2)°, V = 3083.7(1) Å3, and Dc = 4.043 g cm—3 for 3 . 2 is triclinic, space group P1¯, with a = 12.121(2), b = 12.426(3), c = 13.247(3)Å, α = 68.33(3), β = 71.33(3), γ = 71.44(3)°, V = 1710.7(6)Å3 and Dc = 3.735 g cm—3. In all cases, the structures are based on paradodecatungstate polyoxoanions, which are linked by iron ions into chains, layers and a three‐dimensional structure for 1 , 2 and 3 , respectively.  相似文献   

20.
This paper reports the results of the chemical composition modeling for an atmospheric pressure DC air discharge with water cathode. The modeling was based on the combined solution of Boltzmann equation for electrons, equations of vibrational kinetics for ground states of N2, O2, H2O and NO molecules, equations of chemical kinetics and plasma conductivity equation. Calculations were carried out using experimental values of E/N and gas temperatures for the discharge currents range of 20–50 mA. The effect of H2O concentration on the plasma composition was studied. The main particles of plasma were shown to be O2(a1Δ, b1Σ), O(3P), NO, NO2, HNO3, H2O2 and OH. Effective vibrational temperatures of molecules were higher than gas temperature and they did not depend on the discharge current. Distribution functions on vibrational levels for N2, O2, H2O and NO ground states were non-equilibrium ones.  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