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 171 毫秒
1.
The effects of gas composition on hybrid gas–water gliding arc discharge plasma reactor have been studied. The voltage cycles are characterized by a moderate increase in the tension which is represented by a peak followed by an abrupt decrease and a current peak in the half period (10 ms). Emission spectrum measurements revealed that OH hydroxyl radicals are present in the discharge with feeding any gas. The H2O2 concentrations reach 38.0, 15.0, 10.0, and 8.0 mg/l after 25 min plasma treatment with oxygen, argon, air, and nitrogen, respectively. O3 was produced when oxygen and air are used, but not when nitrogen and argon. The O3 concentration reached the highest value 1.0 mg/l after 25 min plasma treatment with oxygen feeding gas, but gradually decreased to 0.2 mg/l after that. With feeding nitrogenous gas, NO2 and NO3 byproducts were formed by the plasma chemical process.  相似文献   

2.

This study investigated the reactive dissolution of nitric oxide (NO) and nitrogen dioxide (NO2) mixtures in deionized water. The dissolution study was carried out in a flat surface type gas–liquid reaction chamber utilizing a gas flow-pattern resembling plasma jets which are often used in biomedical applications. The concentration of NO and NO2 in the gas mixtures was varied in a broad range by oxidizing up to 800 ppm of nitric oxide in Ar carrier gas with variable amount of ozone. The production of nitrite (NO2?) and nitrate (NO3?) in the water was proportional to treatment time up to 50 min. The concentration of NO3? was a power function of gas phase NO2 while the concentration of NO2? increased approximately linearly with gas phase NO2. The formation of NO2? and NO3? could be described by reactions between dissolved NO2 and NO in the water while the production rate was determined by diffusion-limited mass transport of nitrogen oxides to the bulk of the liquid. At higher NO2 concentrations, the formation of dinitrogen tetraoxide (N2O4) increased the formation rate of NO2? and NO3?. The identified mass transport limitation by diffusion suggests that convection of water created by the gas jet is insufficient and dissolution of nitrogen oxides can be increased by additional mixing. In respect of practical applications, the ratio of NO2? /NO3? in water could be varied from 0.8 to 5.3 with treatment time and gas phase NO2 and NO concentrations.

  相似文献   

3.
In this study, hydrogen production from water splitting in N2 using an atmospheric pressure rotating gliding arc plasma was investigated. The effect of input H2O concentration and total flow rate on the performance of the plasma water splitting process (e.g., H2 and O2 yield, H2 production rate, and energy yield of H2) was investigated. N2 showed a pronouncedly facilitating effect on the H2O splitting and H2 production process due to the reactions of the excited N2 species [e.g., electronically excited metastable N2(A)] with the H2O molecules. The maximum H2 production rate reached up to 41.3 μmols?1, which is much higher than that of other typical non-thermal plasmas (e.g., ~0.2 μmols?1 for a dielectric barrier discharge). Optical emission diagnostics has shown that in addition to the NO, N2, and N2 + that were observed in the pure N2 spectra, strong OH and NH emission lines also appeared in the H2O/N2 spectra. OH radical is considered as a key intermediate species that could contribute to the formation of H2, O2, and H2O2. The increase of the H2O concentration could lead to a continuous enhancement of the OH intensity. The rotational temperature of N2 + dropped drastically from 2875 ± 125 to 1725 ± 25 K with the addition of 1 % (mol/mol) H2O into the N2 plasma.  相似文献   

4.
Pulsed arc electrohydraulic discharge (PAED) direct plasma technique was applied for various types of contaminated water treatment. The experimental system consists of a spark-gap switch type pulse power supply (0.5 kJ/pulse) and a 3 L stainless steel reactor with eccentrically configured rod-to-rod electrodes. The current and voltage waveforms are fundamentally different for different conductivity water. Double pulse current discharges were observed for pond water with relatively higher conductivity (637 mS/m) while a single pulse current discharge was observed for lake water with relatively lower conductivity (78 mS/m). From the optical emission spectrum and UV dosimeters, UV-A, UV-B, excited molecules and radicals including N2*, O, H, OH, O3 etc. were observed during the discharge period. Both optical emission and UV intensities in pond water are slightly lower than lake water. The decay time of the UV-A, N2*, OH, H and O radicals were around 0.6 ms, where the discharge period ended around 0.4 ms. The results indicate that the radicals existed longer than the discharge period. The pH, dissolved oxygen and conductivity were changed during the course of PAED treatment. The ions and radicals such as H·, O·, H+, OH· etc. generated by PAED may cause alterations density decay time and the active species in water were present for a longer period. The reduction of total organic carbon (TOC) in pond water reached 80% after 5 min of PAED treatment. Based on local thermal plasma equilibrium (LTE) model, LTE thermal plasma chemical composition model for 1 mol water vapour was used to compare to present experiments. PAED discharge in contaminated pond water generated peak concentration of OH = 28% and O2H = 0.012% mol in gas-phase and migrated to water-phase via gas–liquid interfaces generated by arc and micro-bubbles to form more stable O3, H2O2, H2 and O2. The model suggested that the reduction of TOC occurred in gas/plasma phases or liquid-phase side of gas liquid interfaces.  相似文献   

5.
Comparison of the structures of strychninium N‐phthaloyl‐β‐alaninate N‐phthaloyl‐β‐alanine, C21H23N2O2+·C11H8NO4·C11H9NO4, and brucinium N‐phthaloyl‐β‐alaninate 5.67‐hydrate, C23H27N2O4+·C11H8NO4·5.67H2O, reveals that, unlike strychninium cations, brucinium cations display a tendency to produce stacking inter­actions with cocrystallizing guests.  相似文献   

6.
An infrared spectroscopic study of the diatomic molecules O2, N2, NO and H2 adsorbed under different conditions on Fe2O3 has been performed.Complex patterns of absorption on both α-Fe2O3 and γ-Fe2O3 activated in O2 at high temperature are assigned to vibrations of two different chemisorbed O2 species.N2 molecules do not interact with “oxygen rich” α-Fe2O3 surfaces, but give N2O? and N2O22? species when chemisorbed on evacuated surfaces.NO molecules give complex patterns of absorption, depending on the gas pressure. Three different types of nitrate structures can be identified, as well as NO, NO? and cis-N2O2 chemisorbed species. Chemisorbed water molecules are formed by contact of H2 with Fe2O3 surfaces even at room temperature.  相似文献   

7.
The title compounds are proton‐transfer compounds of cytosine with nicotinic acid [systematic name: 4‐amino‐2‐oxo‐2,3‐dihydropyrimidin‐1‐ium nicotinate monohydrate (cytosinium nicotinate hydrate), C4H6N3O+·C6H4NO2·H2O, (I)] and isonicotinic acid [systematic name: 4‐amino‐2‐oxo‐2,3‐dihydropyrimidin‐1‐ium isonicotinate–4‐aminopyrimidin‐2(1H)‐one–water (1/1/2) (cytosinium isonicotinate cytosine dihydrate), C4H6N3O+·C6H4NO2·C4H5N3O·2H2O, (II)]. In (I), the cation and anion are interlinked by N—H...O hydrogen bonding to form a one‐dimensional tape. These tapes are linked through water molecules to form discrete double sheets. In (II), the cytosinium–cytosine base pairs are connected by triple hydrogen bonds, leading to one‐dimensional polymeric ribbons. These ribbons are further interconnected via nicotinate–water and water–water hydrogen bonding, resulting in an overall three‐dimensional network.  相似文献   

8.
In isomorphous crystals of brucinium 4‐nitro­benzoate methanol solvate, C23H27N2O4+·C7H4NO4·CH3OH, and brucinium 4‐nitro­benzoate dihydrate, C23H27N2O4+·C7H4NO4·2H2O, the brucinium cations form reverse corrugated layers, in which the amine N and amide O atoms of the brucinium cations are located in the grooves and at convex points of the layer surface, respectively. Similarly, as observed for the commonly occurring corrugated brucinium layers, the amide O atoms of the cations are involved in hydrogen bonds in which solvent mol­ecules are the donors.  相似文献   

9.
A strong ionization dielectric barrier discharge was used to produce a high concentration of reactive oxygen species that were then injected into a simulated flue gas in a duct to remove SO2 by oxidation. Sulfuric acid (H2SO4) was produced through the following two reactions: (1) O3 oxidation of SO2–SO3, which then reacted with H2O to produce H2SO4; and (2) reaction of O2 + with H2O to produce ·OH radicals, which then rapidly and non-selectively oxidized SO2–H2SO4. When the molar ratio of reactive oxygen species to SO2 was 4:1, the SO2 removal efficiency was 94.6%, the energy consumption per cubic meter of flue gas was 13.3 Wh/m3, the concentration of recovered H2SO4 was 4.53 g/l, and the H2SO4 recovery efficiency was 28.8%. The H2O volume fraction in the simulated flue gas affected the SO2 removal efficiency, whereas the O2 and CO2 volume fractions did not. These results prove that oxidation by reactive oxygen species is a feasible method for flue gas desulfurization.  相似文献   

10.
By utilizing a fully floating double electrical probe system, the conductivity of a linear atmospheric pressure plasma jet, utilizing nitrogen as process gas, was measured. The floating probe makes it possible to measure currents in the nanoamp range, in an environment where capacitive coupling of the probes to the powered electrodes is on the order of several kilovolts. Using a chemical kinetic model, the production of reactive nitrogen oxide and hydrogen-containing species through admixture of ambient humid air is determined and compared to the measured gas conductivity. The chemical kinetic model predicts an enhanced diffusion coefficient for admixture of O2 and H2O from ambient air of 2.7 cm2 s?1, compared to a literature value of 0.21 cm2 s?1, which is attributed to rapid mixing between the plasma jets and the surrounding air. The dominant charge carriers contributing to the conductivity, aside from electrons, are NO+, NO2 ? and NO3 ?. Upon admixture of O2 and H2O, the dominant neutral products formed in the N2 plasma jet are O, NO and N2O, while O2(1Δg) singlet oxygen is the only dominant excited species.  相似文献   

11.
The atmospheric role of photochemical processes involving NO2 beyond its dissociation limit (398 nm) is controversial. Recent experiments have confirmed that excited NO2* beyond 420 nm reacts with water according to NO2*+H2O→HONO+OH. However, the estimated kinetic constant for this process in the gas phase is quite small (k≈10−15–3.4×10−14 cm3 molecule−1 s−1) suggesting minor atmospheric implications of the formed radicals. In this work, ab initio molecular dynamics simulations of NO2 adsorbed at the air–water interface reveal that the OH production rate increases by about 2 orders of magnitude with respect to gas phase, attaining ozone reference values for NO2 concentrations corresponding to slightly polluted rural areas. This finding substantiates the argument that chemistry on clouds can be an additional source of OH radicals in the troposphere and suggests directions for future laboratory experimental studies.  相似文献   

12.
The chemical interaction between non-thermal plasma species and aqueous solutions is considered in the case of discharges in humid air burning over aqueous solutions with emphasis on the oxidizing and acidic effects resulting from formed peroxynitrite ONOO? and derived species, such as transient nitrite and stable HNO3. The oxidizing properties are mainly attributed to the systems ONOO?/ONOOH [E°(ONOOH/NO2) = 2.05 V/SHE], ·OH/H2O [E°(·OH/H2O) = 2.38 V/SHE] and to the matching dimer system H2O2/H2O [E°(H2O2/H2O) = 1.68 V/SHE]. ONOOH tentatively splits into reactive species, e.g., nitronium NO+ and nitrosonium NO 2 + cations. NO+ which also results from both ionization of ·NO and the presence of HNO2 in acidic medium, is involved in the amine diazotation/nitrosation degradation processes. NO 2 + requires a sensibly higher energy than NO+ to form and is considered with the nitration and the degradation of aromatic molecules. Such chemical properties are especially important for organic waste degradation and bacterial inactivation. The kinetic aspect is also considered as an immediate consequence of exposing an aqueous container to the discharge. The relevant chemical effects in the liquid result from direct and delayed exposure conditions. The so called delayed conditions involve both post-discharge (after switching off the discharge) and plasma activated water. An electrochemical model is proposed with special interest devoted to the chemical mechanism of bacterial inactivation under direct or delayed plasma conditions.  相似文献   

13.
There are many problems with flue gas desulfurization by traditional gas ionization discharge, including the large size of the plasma source, high energy consumption, and the need for a traditional desulfurization method. This paper introduces oxidization of SO2 to sulfuric acid (H2SO4) in a duct by reactive oxygen species (O2 +, O3) produced by strong ionization dielectric barrier discharge. The entire plasma reaction process is completed within the duct without the use of absorbents, catalysts, or large plasma source. The reactive oxygen species O2 + reacts with gaseous H2O in the flue gas to generate ·OH radicals, which can oxidize trace amounts of SO2 in large volumes of the flue gas to produce H2SO4. Sulfuric acid is also produced by O3 oxidation of SO2 to SO3, and SO3 reacting with gaseous H2O in the flue gas. Experimental results showed that with a gas temperature of 22 °C and reactive oxygen species injection rate of 0.84 mg/L, the SO2 removal rate was 81.4 %, and the SO4 2? concentration in the recovered liquid H2SO4 reached 53.8 g/L.  相似文献   

14.
In this paper, we report the effects of the pulse polarity on the plasma jet’s discharge characteristics, particularly, on the production of the reactive oxygen and nitrogen species (RONS) and the inactivation efficiency of myeloma cells, for the purpose of identifying and elucidating the correlation between the dose of RONS and cell viability. Experimental results reveal that the positive plasma jet has a longer length than that for negative plasma jet with the equivalent pulse power. The positive pulse plasma jet would produce higher production of the excited reactive species (OH(A), N2(C), N2+(B), He(3s3S), O(3p5P)), the positive ions (N+, O+, N2+, O2+), and the aqueous species O2?, OH, and ONOO?, while negative plasma jet would generate higher concentration of the negative ions (OH?, O2?, NO2?, NO2?) and the aqueous species NO2? and NO3?. Additionally, the myeloma cells treated by positive plasma jet results in more cell apoptosis and more CD95 expression compared to negative plasma jet, indicating the impact for the cell apoptosis is more significant in the cellular response to the positive plasma jet. By comparing and analyzing the different doses of RONS to the responses of myeloma cells under positive and negative pulse plasma jet, our findings suggest the cell viability has a positive correlation with the concentration of the concentration of ONOO? and the concentration ratio of H2O2 to NO2?, implying the high concentrations for ONOO? and H2O2 might be responsible for the inactivation of myeloma cancer cells.  相似文献   

15.
Crystals of hypoxanthinium (6‐oxo‐1H,7H‐purin‐9‐ium) nitrate hydrates were investigated by means of X‐ray diffraction at different temperatures. The data for hypoxanthinium nitrate monohydrate (C5H5N4O+·NO3?·H2O, Hx1 ) were collected at 20, 105 and 285 K. The room‐temperature phase was reported previously [Schmalle et al. (1990). Acta Cryst. C 46 , 340–342] and the low‐temperature phase has not been investigated yet. The structure underwent a phase transition, which resulted in a change of space group from Pmnb to P21/n at lower temperature and subsequently in nonmerohedral twinning. The structure of hypoxanthinium dinitrate trihydrate (H3O+·C5H5N4O+·2NO3?·2H2O, Hx2 ) was determined at 20 and 100 K, and also has not been reported previously. The Hx2 structure consists of two types of layers: the `hypoxanthinium nitrate monohydrate' layers (HX) observed in Hx1 and layers of Zundel complex H3O+·H2O interacting with nitrate anions (OX). The crystal can be considered as a solid solution of two salts, i.e. hypoxanthinium nitrate monohydrate, C5H5N4O+·NO3?·H2O, and oxonium nitrate monohydrate, H3O+(H2O)·NO3?.  相似文献   

16.
Crystals of brucinium 3,5‐dinitro­benzoate methanol solvate, C23H27N2O4+·C7H3N2O6·CH3OH, (I), brucinium 3,5‐dinitro­benzoate methanol disolvate, C23H27N2O4+·C7H3N2O6·2CH3OH, (II), and brucinium 3,5‐dinitro­benzoate trihydrate, C23H27N2O4+·C7H3N2O6·3H2O, (III), were obtained from methanol [for (I) and (II)] or ethanol solutions [for (III)]. The brucinium cations and 3,5‐dinitro­benzoate anions are linked by ionic N—H+⋯O hydrogen bonds. In the crystals of (I), (II) and (III), the brucinium cations exhibit different modes of packing, viz. corrugated ribbons, pillars and corrugated monolayer sheets, respectively. While in (III), the amide O atom of the brucinium cation participates in O—H⋯O hydrogen bonds, in which water mol­ecules are the donors, in (I) and (II), the amide O atom of the brucinium cation is involved in weak C—H⋯O hydrogen bonds and other brucinium cations are the donors.  相似文献   

17.
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   

18.
Plasma–liquid interactions have gained escalated interests over the last decade due to their potentials in many applications. The simultaneous generation of physicochemical phenomena of interest promotes itself to the top of the promising technologies for liquid processing. Here, we study the physics of a microwave plasma jet (MWPJ) submerged into water and its feasibility to wastewater treatment. We investigate the plasma and bubble dynamics using high-speed imaging. The effects of the argon flow rate, additive gas, and microwave power on the dynamics are examined highlighting the retreating behaviors of plasma channels due to the losses of electrons and power caused by nearby water surface. The addition of N2 (<?5%) to Ar flow results in an oscillatory motion of the foremost edge of the plasma channel. We characterize the submerged MWPJ using a time- and space-averaged optical emission spectroscopy. We found the dominant OH (A–X) molecular band and atomic Ar lines with pure Ar flow indicating the effective dissociation of water. Meanwhile, the addition of N2 leads to an intense emission of NH (A–X) molecular band. Finally, we assess the submerged MWPJ as a viable method for water purification based on the degradation of methylene blue (popular model compound). We find a significant improvement in the efficiency by adding 1–3% of N2 to the Ar, which should be attributed to a combined effects of NH radicals, having high redox potential, and the backward reactions of H2O2 to form OH radicals with NO and NO2.  相似文献   

19.
We have carried out a selected ion flow tube (SIFT) study of the reactions of H3O+, NO+, and O2+ ions with several saturated and unsaturated aldehydes. This study is mainly directed toward providing the essential data for a projected SIFT mass spectrometry (SIFTMS) study of the volatile emissions from cooked meats, which always include aldehydes. Thus, it is necessary to know the rate coefficients and the product ions of the reactions of the above-mentioned ions, used as the precursor ions for SIFTMS analyses, with the aldehydes, if proper identification and quantification of the emitted species are to be achieved. The results of this study show that the reactions of H3O+ with the aldehydes, M, result in the protonated molecules MH+ and for the saturated aldehydes also in (M - OH)+ ions resulting from the loss of a H2O molecule from the nascent MH+ ion. The NO+ reactions invariably proceed via the process of hydride ion, H, transfer producing (M - H)+ ions, but parallel minor association product ions NO+ · M are observed for some of the unsaturated aldehyde reactions. The O2+ reactions proceed by way of charge transfer producing nascent M+ ions that partially dissociate producing fragment ions. Because water vapour is invariably present in real samples analysed by SIFTMS, the current experiments were also carried out following the introduction of humid laboratory air into the helium carrier gas of the SIFT. Thus, the reactions of the product ions that form hydrates were also studied as a prelude to future SIFTMS studies of the (humid) emissions from cooked meats.  相似文献   

20.
The structures of the 1:1 proton‐transfer compounds of 4,5‐dichlorophthalic acid with 8‐hydroxyquinoline, 8‐aminoquinoline and quinoline‐2‐carboxylic acid (quinaldic acid), namely anhydrous 8‐hydroxyquinolinium 2‐carboxy‐4,5‐dichlorobenzoate, C9H8NO+·C8H3Cl2O4, (I), 8‐aminoquinolinium 2‐carboxy‐4,5‐dichlorobenzoate, C9H9N2+·C8H3Cl2O4, (II), and the adduct hydrate 2‐carboxyquinolinium 2‐carboxy‐4,5‐dichlorobenzoate quinolinium‐2‐carboxylate monohydrate, C10H8NO2+·C8H3Cl2O4·C10H7NO2·H2O, (III), have been determined at 130 K. Compounds (I) and (II) are isomorphous and all three compounds have one‐dimensional hydrogen‐bonded chain structures, formed in (I) through O—H...Ocarboxyl extensions and in (II) through N+—H...Ocarboxyl extensions of cation–anion pairs. In (III), a hydrogen‐bonded cyclic R22(10) pseudo‐dimer unit comprising a protonated quinaldic acid cation and a zwitterionic quinaldic acid adduct molecule is found and is propagated through carboxylic acid O—H...Ocarboxyl and water O—H...Ocarboxyl interactions. In both (I) and (II), there are also cation–anion aromatic ring π–π associations. This work further illustrates the utility of both hydrogen phthalate anions and interactive‐group‐substituted quinoline cations in the formation of low‐dimensional hydrogen‐bonded structures.  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