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 33 毫秒
1.
By using spatially resolved Raman spectroscopy together with in situ microscopy, mixed Na2SO4/MgSO4 aerosol particles with molar ratios of 1:1 and 2:1 deposited on a quartz substrate were carefully examined in their evaporation processes. Upon decreasing the relative humidity (RH), phase separations were found to occur for these droplets. For Na2SO4/MgSO4 droplets with a molar ratio of 1:1, two paths of phase separation were identified, and a large amount of small crystals were observed to disorderly distribute in the residual solutions. By comparisons with the known Raman spectra of crystals, it was concluded that the scattered crystals in the two paths were anhydrous Na2SO4 in metastable phase III and the double salt of Na2SO4· MgSO4 · 4H2O, respectively. For Na2SO4/MgSO4 aerosol droplets with a molar ratio of 2:1, only anhydrous Na2SO4 in metastable phase III precipitated with the decrease of RH. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

2.
Raman spectra of potassium, sodium, and ammonium sulfates (K2SO4, Na2SO4, and (NH4)2SO4) are reported and analyzed. These sulfates have been investigated under two states: solid (anhydrous and hydrated) salts and aqueous solutions. The effects of monovalent ions (K+, Na+, and NH4+) and hydration on the position of Raman lines assigned to internal vibrations of sulfate anion SO42− are discussed. In solid salts, the line position of each Raman peak is shown to decrease with increasing radius of the cation. The main ν1 mode of sulfate molecule is particularly affected. It is emphasized that this sensitivity in solid sulfates vanishes in aqueous solutions. As a consequence, this mode can be probed by Raman spectroscopy as the main signature of SO42− to determine its concentration within a single calibration. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

3.
The isotopic composition of air-borne sulphur was investigated in Saxony, Southeast Germany – a region with formerly very high atmospheric SO2 concentrations. In addition, data from various authors were compiled for different Saxonian locations, spanning from 1992 to 2004, i.e., a time of decreasing SO2 concentrations in the atmosphere. There were no obvious temporal changes in the mean δ34S value of bulk precipitation. However, the variability of monthly mean δ34S values decreased. The mean sulphur isotope composition of sulphate from bulk precipitation after the year 2000 converges in Saxony towards 4–5‰, with similar values for different locations. Mean values of different forms of sulphur show the following enrichment order: δ34S of SO2 < δ34S of weathering crusts ≤δ34S of sulphate from bulk precipitation ≤δ34S of dust. Judging from local differences on sulphate crusts and corresponding isotope values of sources, the δ34S value of SO2 as well as for crusts mainly reflects local point sources. The mean δ34S value of bulk precipitation represents more regionally well-mixed SO2 sources and is therefore an ideal tool for monitoring regional atmospheric change.  相似文献   

4.
The use of surface enhanced Raman scattering (SERS) to study oxidation-reduction and complexation chemical reactions on Au surfaces is illustrated by: (1) the reaction of Au(CN)2? adsorbed on a Au colloid (2140 cm?) to form Au(CN)32? (2131 cm?1) on the surface in excess CN?; (2) the oxidation of Au(CN)2? by HNO3, Cl2, or Br2 solutions to form Au(CN)4? (2190 cm?) on a Au colloid; and (3) the dissolution of Au in excess CN? with O2. Unlike with Ag surfaces, no SERS is observed when Au powder is exposed to NO, NO2, SO2, CO, or CO2 gases. The surface chemistry of Au is discussed in the light of these reactions.  相似文献   

5.
《Solid State Ionics》1986,20(1):61-68
Electrical conductivity data are reported for solid solutions of Na2SO4, K2WO4, Na2WO4, Na2MoO4, Rb2SO4, Na4SiO4 and Gd2 (SO4)3. In all cases, except K2SO4, we observed an increase in Na+ conductivity effected by lattice expansion and/or incorporation of ion vacancies in addition to a structural transformation. Boundary conditions were shown to exist for these factors to yield a limiting Na+ conductivity with a constant fraction of Na+ based on a percolation model of transport. The higher conductivity data observed for the larger radius isovalent WO2-4 and aliovalent SiO4-4 doped Na2SO4 show conclusively that the anion-rotation ”cogwheel” mechanism does not contribute to the cationic conductivity in Na2SO4.  相似文献   

6.
Sonochemical species such as nitrite (NO2) and nitrate (NO3) were detected in ultrapure aqueous medium with 28 kHz low frequency ultrasound (US) in the range of 200–1200 W output power. The concentration of their anionic ions monitored with a high-performance liquid chromatography increased with increasing US power especially under air atmosphere. When the generation of NO2 and NO3 ions under US exposure was investigated for N2, O2 and Ar-bubbled solutions, no trace of NO2 was observed while NO3 was slightly generated. Under air atmosphere, the concentration of dissolved oxygen in the aqueous medium increased especially when 1200 W power was used. In addition, the bulk pH shifted towards the acidic side with an increase in exposure time, which indicated that NO2 was formed. The formation of oxidizing species under 28 kHz low frequency ultrasonic treatment was confirmed with an absorption spectrum which was dominated by two maxima peaks at 288 nm and 352 nm representing triiodide I3 anion.  相似文献   

7.
The selective catalytic reduction (SCR) of nitrogen dioxide in an air flow modeling the exhaust gas from an internal combustion engine is studied. Granulated V2O5 (13.5%)–MnO2 (0.7–1.0%)/Al2O3 powder (AVK-10M catalyst) and ammonia injected into a SCR catalytic cell are used as a heterogeneous catalyst of NO2 reduction and a reducing agent, respectively. If the efficiency of NO2 removal is high enough and satisfies the requirements of the State Sanitary Standards for the maximum permissible concentrations of substances emitted into the atmosphere (\(MP{C_{N{O_2}}}\) = 0.085 mg/m3), the reducing agent (ammonia) is not completely consumed during SCR, so a considerable amount of NH3 can be released into the atmosphere. Therefore, a strict control of both NO2 and unreacted ammonia emissions is needed. The dependences of the concentrations of [NH3] and [NO2] on the [NH3]/[NO2] ratio for the model air flow passed through the AVK-10M granular heterogeneous catalyst are measured. It is found that the maximum degree of removal of NO2 from the air takes place at [NH3]/[NO2] = 1.3. In the conventional process, the concentration of [NO2] drop from 530.00 to 0.07 mg/m3, i.e., below the \(MP{C_{N{O_2}}}\). At the same time, the ammonia concentration increases to [NH3] = 3.4 mg/m3, which becomes 85 times the \(MP{C_{N{O_2}}}\), 0.04 mg/m3. To remove unreacted ammonia from air flows, we developed [P–(SO3 -)2 · Me2+] sulfocationites, where Me are the Cu and/or Ca ions, P is a styrene–divinylbenzene copolymer. It is shown that the concentration of ammonia passed through the adsorption cell filled with a freshly sulfocationite drops below \(MP{C_{N{H_3}}}\) = 0.04 mg/m3. The dependences of the dynamic exchange capacity (DEC) before ammonia breakthrough for the [P–(SO3 -)2 · Cu2+] delta-sulfocationite on the air flow rate, [NH3] concentration, and humidity are measured. The maximum value of the DEC, δ = 59.5 mg/cm3, is observed at an air flow velocity of 2.171 m/s, [NH3] = 0.0035 mg/L, and 75% humidity. To illustrate practical applications of the proposed improved SCR method, it is shown that a 3-L replaceable [P–(SO3 -)2 · Cu2+] sorbent cartridge in a SCR exhaust gas purifier for a car internal combustion engine does not need replacement more frequently than every 50000 km.  相似文献   

8.
《Solid State Ionics》1987,25(1):41-44
Complexes of alkali metal salts with various polymers have for some time been recognized as fast ionic conductors. Polymer electrolyte fast ion conductors are currently under consideration for use in high energy density electrochemical cells. In order to aid in our understanding of the mechanism of ionic conductivity we have examined systematically complexes of poly(ethylene oxide) (PEO) with the alkali metal salt series of Li+, Na+, K+, Rb+ and Cs+ with both tetraflouroborate (BF4-) and trifluoromethanesulfonate (CF3SO4-) anions. The ratio of monomer to salt was 10:1 in all cases. Complex impedance measurements were made on all samples in the temperature range 40°–125°C. With CF3SO4- as the anion a definite trend was apparent with the smallest cation Li+ being the worst conductor and Cs+, the largest cation, being the best. When BF4- salts are used, the Na+ complex is found to be the best conductor and Rb+ the worst. This study, in connection with our earlier studies, has shown that synergy between cation and anion in the polymer matrix is an important consideration in determining the ionic conductivity.  相似文献   

9.
In the binary system (1?x)Li2SO4xNa2SO4, the solid–solid phase transitions and energy storage properties of Li2SO4, Na2SO4, the binary compound LiNaSO4 and two eutectoids (E1: 0.726Li2SO4–0.274Na2SO4; E2: 0.03Li2SO4–0.97Na2SO4) were investigated by X-ray diffraction and differential scanning calorimetry. Li2SO4 has a solid–solid phase transition at 578 °C with the transition enthalpy 252 J g?1. The binary compound LiNaSO4 gives a slightly lower enthalpy value, 214 J g?1 and its transition temperature is clearly reduced to 514 °C. The transition enthalpy of the eutectoid E1 is maintained to 177 J g?1 and its transition temperature is further reduced to 474 °C. Li2SO4, LiNaSO4 and the eutectoid E1 are applicable phase transition materials because of their large transition enthalpies. The enthalpies of Na2SO4 and the eutectoid E2 are not very high (~45 J g?1), but their transition temperatures are quite low (~250 °C); thus their transition properties may be applied at such low temperatures.  相似文献   

10.
A flow hydrothermal setup with a tubular reactor equipped with a plunger pump and back pressure valves is used to study the effects of scaling in the K2SO4-KCl-H2O, K2SO4-K2CO3-H2O, and Na2SO4-NaCl-H2O systems at pressures of up to 270–340 kg/cm2, temperatures of 400–600°C, and flow rates of 5.0 and 2.5 ml/min in order to establish conditions for the formation of salt plugs of type 2 (K2SO4, Na2SO4) in the flow mode at supercritical (SC) state parameters and to explore ways of eliminating such salt deposits by means of hydrothermal solvents, more specifically, high-temperature aqueous solutions of salts of type 1 (KCl, K2CO3, and NaCl). The concentrations of hydrothermal solvents sufficient to prevent the plugging of flow systems with solutions containing 0.26–0.27 mol % K2SO4 or Na2SO4 are determined, and the effects of the flow rate and chemical composition of type 1 salts on this process are studied. The results show that the phenomenon of scaling with the formation of salt plugs, which hinders the practical use of supercritical water oxidation technology, can be eliminated by adding readily soluble electrolytes, salts of type 1, to initial aqueous solution of type 2 salts.  相似文献   

11.
The effects of Cl, NO3 and SO42− aggressive anions on the corrosion and passivation behavior of carbon steel electrode in deaerated 0.50 M NaHCO3 solutions were studied using potentiodynamic anodic polarization and SEM techniques. It was found that the presence of Cl, NO3 and SO42− anions stimulates the anodic dissolution rate in both the active and the pre-passive potential regions. Moreover, significantly great effects were observed in both the passive and the trans-passive potential regions. Pitting corrosion was observed only in the presence of Cl anions, while the presence of NO3 and SO42− anions facilitate only passivation by oxygen of water without themselves participating in the cathodic process. Also, it was observed that the effect of NO3 anion, which is a strong oxidizing agent acting “primarily” as stimulator of the cathodic process and then its reaction product acts “indirectly” retarding the anodic process. On the other hand, the effect of SO42− anion, which is a non-oxidizing agent, exerts an “indirect” effect on the cathodic reaction increasing its rate and then “directly” influence on the anodic reaction, retarding it.  相似文献   

12.
It has been shown recently that when a relatively weak absorber is placed within a laser cavity an enhacement of absorption occurs1–3. This method has been succesfully used for trace analysis of Na1,4, I2 3,5, Sr and Ba+ 6, Eu(NO3)3 2,8, Pr(NO3)3, NdCl3 and HoCl3.8 In these experiments the absorbing species were placed inside the cavity of: flashlamp-pumped dye lasers1–4,6 continous wave dye lasers3,5 and dye lasers pumped by a ruby laser8.  相似文献   

13.
李立曼  王刚 《物理学报》1989,38(5):849-852
本工作利用X射线衍射分析、电导测量、红外、喇曼散射等手段,对用熔盐电解法生长的KxWO3,NayWO3和KxNayWO3单晶进行了深入的研究,得出KxNayWO3是钠离子插入到KxWO3的结论,提出了KxNay关键词:  相似文献   

14.
The processes of molecular relaxation in the solid NaNO3–NaNO2 and KNO3–KNO2 “nitrate–nitrite” binary systems have been investigated by Raman spectroscopy. The relaxation time of the vibration ν1(A) of an NO- 3 anion in the binary system is found to be shorter than that in individual nitrate. The increase in the relaxation rate is explained by the existence of an additional mechanism of relaxation of vibrationally excited states of the nitrate ion in the system. This mechanism is related to the excitation of vibration of another anion (NO- 2) and generation of a lattice phonon. It has been established that this relaxation mechanism is implemented provided that the difference between the frequencies of the aforementioned vibrations correspond to the range of sufficiently high density of states in the phonon spectrum.  相似文献   

15.
《Solid State Ionics》1987,23(3):151-163
Premelted, predried Na2SO4, premelted Na2WO4, Na2SO4Na2WO4 composites and Na2SO4M2(SO4)3 (M = La, Dy, Sm, In) have been studied by means of X-ray diffraction, DTA and electrical conductivity measurements. The high temperature, highly conducting Na2SO4 phase I has been successfully stabilised at room temperature; the Na2SO4 containing 4 mole% La2(SO4)3 exhibits the highest conductivity (σ) and lowest activation energy (Ea) (σ=1.08 × 10−3ω−1 cm−1 at 290°C and Ea=0.50 eV) and therefore this system appears promising for further development.  相似文献   

16.
Abstract

The Raman spectrum of mercury(II) iodide was observed in cesium nitrate at 430°C and in meta- and para-terphenyl at 235 and 240°C, respectively. A single strong polarized line was found at 148. 5 cm?1, width at half height 15 cm?1 in cesium nitrate, and a single line at 154. 0 cm?1 width at half height 7 cm?1 in the terphenyls. The observed spectra are not consistent with interactions of the solute and the solvents involving bonds of highly covalent character, but do not exclude other interactions.

In the course of an investigation on the nature of mercury (II) iodide species in solution in molten alkali metal nitrates, by its distribution between the salt melts and terphenyl melts1, 2, it became of interest to study by means of Raman spectroscopy the mercury (II) iodide species formed. Several such studies at lower temperatures have already been made: in the gas phase3, in a krypton matrix4, in alcohols5, tributy phosphate6 and dioxame7, and in molten mercury(II) iodide8, chloride9, and bromide9. The reason for looking at the Raman spectrum in yetfurther media was the suggestion made on the basis of thermodynamic and kinetic data2,10 that the mercury(II) iodide species in the alkali mitrate melts are solvated by nitrate anions, and that possiby the mixed anion terrahedral species HgI2(NO3)2 is formed. Recent Ranan sepctrophotometric data on mixed halide anionic complexes of mercury11 identified prominent lines of the spectrum of the species HgBrnI2– 4–n, including HGBr2I2 2–, so that a comparison could be made. The solubility of mercury (II) iodide in molten alkali metal nitrates is rather small, expect for cesium nitrate12, Where the solubility should be sufficent for the Raman spectrum to be recorded. Also, it was comcluded from vapor pressure osmometric data in aromatic solvents that demiric species of mercury(II) halides (even a trimeric species of the iodide) are found, in which the mercury has a distorted octahedral coordination, with three halogen atoms bonded to a mercury atom in the dimer13. In terphenyl melts, the solubility  相似文献   

17.
Although the amine sulfur dioxide chemistry was well characterized in the past both experimentally and theoretically, no systematic Raman spectroscopic study describes the interaction between N,N‐dimethylaniline (DMA) and sulfur dioxide (SO2). The formation of a deep red oil by the reaction of SO2 with DMA is an evidence of the charge transfer (CT) nature of the DMA–SO2 interaction. The DMA–SO2 normal Raman spectrum shows the appearance of two intense bands at 1110 and 1151 cm−1, which are enhanced when resonance is approached. These bands are assigned to νs(SO2) and ν(ϕ N) vibrational modes, respectively, confirming the interaction between SO2 and the amine via the nitrogen atom. The dimethyl group steric effect favors the interaction of SO2 with the ring π electrons, which gives rise to a π–π* low‐energy CT electronic transition, as confirmed by time‐dependent density functional theory (TDDFT) calculations. In addition, the calculated Raman DMA–SO2 spectrum at the B3LYP/6‐311 + + g(3df,3pd) level shows good agreement with the experimental results (vibrational wavenumbers and relative intensities), allowing a complete assignment of the vibrational modes. A better understanding of the intermolecular interactions in this model system can be extremely useful in designing new materials to absorb, detect, or even quantify SO2. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

18.
A study of the influence of cationic Na+ substitution in the archetype KMnF3 perovskite crystal was performed using the Raman method. The Raman spectra of mixed K1-xNaxMnF3 crystals with x = 0.029, 0.048 and 0.065 were recorded versus temperature and fully interpreted in terms of a “one mode” behaviour. In addition to the soft mode not completely vanishing close to Tc, attention was especially paid to evidence of static and dynamical disorder. From this point of view the behaviour of the hard Raman modes versus temperature has been studied together with two unexpected Raman bands in the cubic phase. The interpretation has been made within the more general framework of structural disorder existing in such perovskites with anisotropic interactions.  相似文献   

19.
Five Na2SO4:RE3+ phosphors activated with rare-earth (RE) ions (RE3+=Ce3+, Sm3+, Tb3+, Dy3+ and Tm3+) were synthesized by heating natural thenardite Na2SO4 from Ai-Ding Salt Lake, Xinjiang, China with small amounts of rare-earth fluorides, CeF3, SmF3, TbF3, DyF3 and TmF3, at 920 °C in air. The photoluminescence (PL) and optical excitation spectra of the obtained phosphors were measured at 300 and 10 K. In the PL spectrum of Na2SO4:Ce3+ at 300 K, two overlapping bands with peaks at 335 and 356 nm due to Ce3+ were first observed. Narrow bands observed in PL and excitation spectra of Na2SO4:RE3+ (RE3+=Sm3+, Tb3+, Dy3+ and Tm3+) phosphors were well identified with the electronic transitions within the 4fn (n=5, 8, 9 and 12) configurations of RE3+. The existence of excitation bands with high luminescence efficiency at wavelengths shorter than 230 nm is characteristic of Na2SO4:RE3+ (RE3+=Sm3+, Tb3+, Dy3+ and Tm3+) phosphors. The obtained results suggest that these phosphors are unfavorable as the phosphor for usual fluorescence tubes, i.e., mercury discharge tubes, but may be favorable as the phosphor for UV-LED fluorescent tubes and as cathodoluminescence, X-ray luminescence and thermoluminescence phosphors.  相似文献   

20.
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