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 718 毫秒
1.
We present density functional theory (DFT) calculations on the formation of nitric oxide dimers (N2O2) on Au atoms, dimers and trimers adsorbed on regular O2 ? sites and neutral oxygen vacancies (Fs sites) of the MgO(100) surface. The study of the N2O2 species is of great interest since it has been detected in the NO reduction reaction as an intermediate towards the formation of N2O. We found that the coupling of a NO molecule with a previously adsorbed one on Au/MgO is energetically favorable on Au1 and Au3, but unfavorable on Au2. The stability of N2O2 is in direct relation with the amount of charge taken from the support. Furthermore, one of the N―O bonds can be activated as a result of the attraction between the negatively charged NO dimer and the ionic oxide surface. In fact, for Au1 anchored on the Fs site a barrierless reaction occurs between N2O2 and a third NO molecule, forming adsorbed N2O and NO2.  相似文献   

2.
Metal-oxide nanocomposites were prepared over screen-printed gold electrodes to be used as room-temperature NOx (nitric-oxide (NO) and nitrogen dioxide (NO2)) sensors. Various weight ratios of SnO2–WO3 and Pt loadings were used for NO sensing. The sensing materials were characterized using X-ray diffraction (XRD), transmission electron microscopy (TEM) and BET surface analysis. The NO-sensing results indicated that SnO2–WO3 (1:2) was more effective than other materials were. The sensor response (S=resistance of N2/resistance of NO=RN2/RNO) for detecting 1000 ppm of NO at room temperature was 2.6. The response time (T90) and recovery time (TR90) was 40 s and 86 s, respectively. By further loading with 0.5% Pt, the sensor response increased to 3.3. The response and recovery times of 0.5% Pt/SnO2–WO3 (1:2) were 40 s and 206 s, respectively. The linearity of the sensor response for a NO concentration range of 10–1000 ppm was 0.9729. A mechanism involving Pt promotion of the SnO2–WO3 heterojunction was proposed for NO adsorption, surface reaction, and adsorbed NO2 desorption.  相似文献   

3.
Following our recent work on IR spectra of molecules adsorbed on C60 embedded in LiF and LiCl films, adsorption potentials of CO and N2 adsorbed on LiF (100) and LiCl (100) were calculated. For CO on LiF, a value of 2.0 kcal mol−1 was obtained, close to that calculated for CO adsorbed on a single C60 molecule. The calculated value for CO on LiCl is much higher, 6.8 kcal mol−1. It is therefore concluded that in the case of CO adsorbed on mixed LiF/C60 films, the adsorbed CO molecules are distributed almost evenly on the LiF and C60 single molecules, whereas in the case of CO adsorbed on mixed LiCl/C60 films the salt is greatly preferred as the adsorption site. Adsorption potential calculations for a similar system, N2 on LiF and LiCl, gave values of 1.5 and 4.4 kcal mol−1, respectively. In this case, a much too large value was found for the adsorption potential on the LiCl surface. IR spectra of CO on the two substrates showed two strong absorptions for each of them. With N2 induced spectra were obtained. Spectral shifts have been calculated for the above systems and were all toward higher frequencies, in agreement with experimental findings.  相似文献   

4.
The adsorption properties of a variety of atomic species (H, O, N, S, and C), molecular species (N2, HCN, CO, NO, and NH3) and molecular fragments (CN, NH2, NH, CH3, CH2, CH, HNO, NOH, and OH) are calculated on the (111) facet of palladium using periodic self-consistent density functional theory (DFT–GGA) calculations at ¼ ML coverage. For each species, we determine the optimal binding geometry and corresponding binding energy. The vibrational frequencies of these adsorbed species are calculated and are found to be in good agreement with experimental values that have been reported in literature. From the binding energies, we calculate potential energy surfaces for the decomposition of NO, CO, N2, NH3, and CH4 on Pd(111), showing that only the decomposition of NO is thermochemically preferred to its molecular desorption.  相似文献   

5.
Nitric oxide chemistry and photochemistry on the Cr-terminated surface of α-Cr2O3(0001) were examined using temperature programmed desorption (TPD), sticking coefficient measurements and photodesorption. NO exposed to α-Cr2O3(0001) at 100 K binds at surface Cr cation sites forming a strongly bound surface species that thermally desorbs at 320–340 K, depending on coverage. No thermal decomposition was detected in TPD in agreement with previous results in the literature. Sticking probability measurements at 100 K indicated near unity sticking for NO up to coverages of ~ 1.3 ML, with additional adsorption with higher exposures at decreased sticking probability. These results suggest that some Cr cation sites on the α-Cr2O3(0001) surface were capable of binding more than one NO molecule, although it is unclear whether this was as separate NO molecules or as dimers. Photodesorption of adsorbed NO was examined for surface coverages below the 1 ML point. Both visible and UV light were shown to photodesorb NO without detectable NO photodecomposition. Visible light photodesorption of NO occurred with a greater cross section than estimated using UV light. The visible light photodesorption event was not associated with bandgap excitation in α-Cr2O3(0001), but instead was linked to excitation of a surface Cr3 +–NO? charge transfer complex. These results illustrate that localized photoabsorption events at surface sites with unique optical properties (relative to the bulk) can result in unexpected surface photochemistry.  相似文献   

6.
By introducing the covalent effect into the Girerd–Journaux–Kahn's magnetic theory a more general magnetic exchange formula for describing transition-metal ion pairs in covalent complex molecules has been established. By this formula and by the use of our double-slater-type-function (DSF) calculation procedure the relationship between the magnetic exchange coupling parameter J and the covalent factor N has been studied. It is shown that for the oxo-bridged Fe(III) dimer there exists a nearly linear relation between J and N, and that the stronger is the covalent effect the stronger is the antiferromagnetic coupling interaction. It is also shown that in a series of model molecules of methemerythrin derivatives, such as in (FeSalen)2O·2py, FeSalenCl·(CH3NO2)x, enH2[(FeHEDTA)2O]·6H2O, Na4[(FeHEDTA)2O]·12H2O, [(Fe(terpy))2O](NO3)4·H2O and Fe(terpy)Cl3, there are about 40% antiferromagnetic contributions that arise from the covalent effect, and that the theoretical values J=−95.5 to −100.7 cm−1 are in good agreement with the experimental findings J=−90 to −105 cm−1 for those model molecules.  相似文献   

7.
《Physics letters. A》2020,384(16):126332
Hydrogen-substituted graphdiyne (HsGDY) is a novel alkynyl carbon material with a structure similar to that of graphene. In this paper, the adsorption of four gas molecules (NO, NO2, NH3, and N2) on HsGDY and B-doped HsGDY (B-HsGDY) was studied using density functional theory. The results show that the adsorption of NO and NO2 on HsGDY and B-HsGDY is characterized by a larger charge transfer, stronger interaction, and higher adsorption energy compared with that of NH3 and N2. Based on the doping with B atoms, the adsorption energies of the gas molecules on HsGDY significantly improve, especially that of NO and NO2. The gas molecule adsorption on both HsGDY and B-HsGDY is physical adsorption and the adsorption selectivity is good and thus may be applied for gas-sensitive NO and NO2 materials.  相似文献   

8.
Adsorption and decomposition of NO on Pt (1 1 2) have been studied by temperature programmed desorption (TPD), ultraviolet photoelectron spectroscopy (UPS) and X-ray photoelectron spectroscopy (XPS). NO adsorbs molecularly on Pt (1 1 2) at 95 K. About half amount of NO molecules adsorbs at the terrace sites and remaining half amount adsorbs at the step sites at a full monolayer coverage. Then about half of NO molecules adsorbed at step sites decomposes at around 483 K desorbing N2, promptly.  相似文献   

9.
Karl Jacobi  Yuemin Wang 《Surface science》2009,603(10-12):1600-1604
The interaction of NO with the O-rich RuO2(1 1 0) surface, exposing coordinatively unsaturated O-bridge, O-cus, and Ru-cus atoms, was studied at 300 K by thermal desorption spectroscopy (TDS) and high-resolution electron energy-loss spectroscopy (HREELS). The conclusions are validated by isotope substitution experiments with 18O. During exposure to NO an O···N–O surface group (NO2-cus) is formed with O-cus. Additionally, a smaller number of empty Ru-cus sites are filled by NO-cus. If one warms the sample to 400 K, NO2-cus does not desorb but decomposes into O and NO again, the latter being either released into gas phase or adsorbed as NO-cus. With O-bridge such a surface group is not stable at 300 K. Our experiments further prove that O-cus is more reactive than O-bridge.  相似文献   

10.
Density-functional calculations of molecular nitric oxide (NO) on defective (La,Sr)O (001) surfaces of (La,Sr)FeO3 ? δ using slab models are performed to elucidate the oxygen vacancy formation problem on the LaO (001) surface of LaFeO3 ? δ.From the estimation of the NO adsorption energy, NO adsorption is found on (La,Sr)O surfaces of (La0.83,Sr0.17)FeO3 ? δ with δ = 0 or 0.25.The absolute value of the NO adsorption energy shows a remarkable increase at oxygen vacancies in the top surface layer, where the nitrogen atoms of the adsorbed molecules are embedded in the first (La,Sr)O layer, because a bond with Fe in the second FeO2 layer is formed.Our data shows that Sr doping promotes formation of oxygen vacancies, which keep the NO adsorption ability high.Thus, we conclude that if Sr doping increases the number of oxygen vacancy sites by a charge compensation effect, NO adsorption on LaFeO3 is enhanced, which provides an explanation for several experimental observations.  相似文献   

11.
Continuous-time photoelectron spectroscopy (PES) and photon-exposure-dependent photon-stimulated desorption (PSD) were employed to investigate the monochromatic soft X-ray-induced dissociation of SF6 molecules adsorbed on Si(111)-7 × 7 at 30 K (SF6 dose = 3.4 × 1013 molecules/cm2, ~ 0.5 monolayer). The photon-induced evolution of adsorbed SF6 was monitored at photon energies of 98 and 120 eV [near the Si(2p) edge], and sequential valence-level PES spectra made it possible to deduce the photolysis cross section as a function of energy. It was found that the photolysis cross sections for 98 and 120 eV photons are ~ 2.7 × 10? 17 and ~ 3.7 × 10?17 cm2, respectively. The changes in the F? and F+ PSD ion yields were also measured during irradiation of 120 eV photons. The photon-exposure dependencies of the F? and F+ ion yields show the characteristics: (a) the dissociation of adsorbed SF6 molecules is ascribable to the substrate-mediated dissociations [dissociative attachment (DA) and dipolar dissociation (DD) induced by the photoelectrons emitting from the silicon substrate]; (b) at early stages of photolysis, the F? yield is mainly due to DA and DD of the adsorbed SF6 molecules, while at high photon exposure the F? formation by electron capture of the F+ ion is likely to be the dominant mechanism; (c) the F+ ion desorption is associated with the bond breaking of the surface SiF species; (d) the surface SiF is formed by reaction of the surface Si atom with the fluorine atom or F? ion produced by scission of S–F bond of SFn (n = 1–6) species.  相似文献   

12.
By applying pulsed high voltage discharge to a needle-mesh reactor that using seven acupuncture needles as discharge electrode and stainless steel wire mesh as ground electrode, nitrogen from bubbling gas could be fixed into NO2? and NO3? with equivalent mol H+ produced in the liquid phase and a small amount of NO and NO2 yielded in the gas phase. The HNO2 was originally formed and then converted into HNO3. The ·OH and H2O2 stimulated the conversion reaction from HNO2 to HNO3, which caused HNO2 concentration increased in the first 12 min and then decreased until lower than its detection limit. The concentration of HNO3 still increased with discharge time. After 36 min, HNO3 was the only and ultimate product in the liquid. The total yield of HNO2 and HNO3 could be affected by processing parameters such as electric factors of peak voltage and frequency, mesh size of ground electrode and content of nitrogen in N2/O2 bubbling. Increasing peak voltage or frequency, the total yield of HNO2 and HNO3 increased. Gas composition had a heavy impact on the fixation efficiency that obtained its maximum value at an oxygen content of 66.7% with bubbling O2/N2 gas. At the end of the 36 min discharge, the HNO3 concentration with bubbling air was 2.215 mmol L?1 at an applied voltage of 25 kV, pulse repetition frequency of 140 Hz and ground electrode mesh of 20 × 20. The energy yield was about 1.22 g (HNO3)/kWh.  相似文献   

13.
Continuous-time photoelectron spectroscopy (PES) and continuous-time core-level photon-stimulated desorption (PSD) spectroscopy were used to study the monochromatic soft X-ray-induced reactions of CCl2F2 molecules adsorbed on Si(111)-7 × 7 at 30 K (CCl2F2 dose = 2.0 × 1014 molecules/cm2, ~ 0.75 monolayer) near the Si(2p) core level. Evolution of adsorbed CCl2F2 molecules was monitored by using continuous-time photoelectron spectroscopy at two photon energies of 98 and 120 eV to deduce the photolysis cross section as a function of energy. It was found that the photolysis cross sections for 98 and 120 eV photons are ~1.4 × 10? 18 and ~ 8.0 × 10? 18 cm2, respectively. Sequential F+ PSD spectra obtained by using continuous-time core-level photon-stimulated desorption spectroscopy in the photon energy range of 98–110 eV show the variation of their shapes with photon exposure and depict the formation of surface SiF species. The dissociation of CCl2F2 molecules adsorbed on Si(111)-7 × 7, irradiated by monochromatic soft X-ray in the photon energy range of 98–110 eV, is mainly due to dissociative electron attachment and indirect dipolar dissociation induced by photoelectrons emitted from the silicon surface.  相似文献   

14.
Chitosan acetate-ammonium nitrate (NH4NO3) films have been prepared by the solution-cast technique. Fourier transform infrared spectroscopy (FTIR) showed that complexation has occurred. FTIR exhibited shifts in amine and carbonyl bands from 1553 to 1520 cm−1 and 1636 to 1617 cm−1. A new peak was also observed at 1746 cm−1. XRD shows that all complexes are amorphous. The highest conductivity at room temperature is 2.53×10−5 S cm−1 for the film containing 45 wt% NH4NO3. The conductivity of the samples is dependent on the number of mobile ions and mobility.  相似文献   

15.
Michael A. Henderson 《Surface science》2010,604(19-20):1800-1807
The photochemical properties of the Cr-terminated α-Cr2O3(0001) surface were explored using methyl bromide (CH3Br) as a probe molecule. CH3Br adsorbed and desorbed molecularly from the Cr-terminated α-Cr2O3(0001) surface without detectable thermal decomposition. Temperature programmed desorption (TPD) revealed a CH3Br desorption state at 240 K for coverages up to 0.5 ML, followed by more weakly bound molecules desorbing at 175 K for coverages up to 1 ML. Multilayer exposures led to desorption at ~ 130 K. The CH3Br sticking coefficient was unity at 105 K for coverages up to monolayer saturation, but decreased as the multilayer formed. In contrast, pre-oxidation of the surface (using an oxygen plasma source) led to capping of surface Cr3+ sites and near complete removal of CH3Br TPD states above 150 K. The photochemistry of chemisorbed CH3Br was explored on the Cr-terminated surface using post-irradiation TPD and photon stimulated desorption (PSD). Irradiation of adsorbed CH3Br with broad band light from a Hg arc lamp resulted in both photodesorption and photodecomposition of the parent molecule at a combined cross section of ~ 10? 22 cm2. Photodissociation of the CH3–Br bond was evidenced by both CH3 detected in PSD and Br atoms left on the surface. Use of a 385 nm cut-off filter effectively shut down the photodissociation pathway but not the parent molecule photodesorption process. From these observations it is inferred that d-to-d transitions in α-Cr2O3, occurring at photon energies < 3 eV, do not significantly promote photodecomposition of adsorbed CH3Br. It is unclear to what extent band-to-band versus direct CH3Br photolysis play in CH3–Br bond dissociation initiated by more energetic photons.  相似文献   

16.
We demonstrated efficient red organic light-emitting diodes based on a wide band gap material 9,10-bis(2-naphthyl)anthracene (ADN) doped with 4-(dicyano-methylene)-2-t-butyle-6-(1,1,7,7-tetramethyl-julolidyl-9-enyl)-4H-pyran (DCJTB) as a red dopant and 2,3,6,7-tetrahydro-1,1,7,7,-tetramethyl-1H,5H,11H-10(2-benzothiazolyl)quinolizine-[9,9a,1gh]coumarin (C545T) as an assistant dopant. The typical device structure was glass substrate/ITO/4,4′,4″-tris(N-3-methylphenyl-N-phenyl-amino)triphenylamine (m-MTDATA)/N,N′-bis(naphthalene-1-yl)-N,N′-diphenyl-benzidine (NPB)/[ADN:Alq3]:DCJTB:C545T/Alq3/LiF/Al. It was found that C545T dopant did not by itself emit but did assist the energy transfer from the host (ADN) to the red emitting dopant via cascade energy transfer mechanism. The OLEDs realized by this approach significantly improved the EL efficiency. We achieved a significant improvement regarding saturated red color when a polar co-host emitter (Alq3) was incorporated in the matrix of [ADN:Alq3]. Since ADN possesses a considerable high electron mobility of 3.1 × 10−4 cm2  V−1 s−1, co-host devices with high concentration of ADN (>70%) exhibited low driving voltage and high current efficiency as compared to the devices without ADN. We obtained a device with a current efficiency of 3.6 cd/A, Commission International d’Eclairage coordinates of [0.618, 0.373] and peak λmax = 620 nm at a current density of 20 mA/cm2. This is a promising way of utilizing wide band gap material as the host to make red OLEDs, which will be useful in improving the electroluminescent performance of devices and simplifying the process of fabricating full color OLEDs.  相似文献   

17.
Improved performance of organic light-emitting diodes (OLEDs) as obtained by a mixed layer was investigated. The OLEDs with a mixed layer which were composed of N,N′-diphenyl-N,N′-bis(1-napthyl-phenyl)-1,1′-biphenyl-4,4′-diamine (NPB), tris-(8-hydroxyquinolato) aluminum (Alq3) and 4-(dicyanomethylene)-2-t-butyl-6-(1,1,7,7-tetramethyljulolidyl-9-enyl)-4H-pyran (DCJTB) showed the highest brightness and efficiency, which reached 19048 cd/m2 at 17 V and 4.3 cd/A at 10 mA/cm2, respectively. The turn-on voltage of the device is 2.6 V. Its Commission Internationale del’Eclairage (CIE) coordinate is (0.497, 0.456) at 17 V, and the CIE coordinates of the device are largely insensitive to the driving voltages, which depicts stabilized yellow color.  相似文献   

18.
《Solid State Ionics》2006,177(37-38):3223-3231
Proton dynamics in (NH4)3H(SO4)2 has been studied by means of 1H solid-state NMR. The 1H magic-angle-spinning (MAS) NMR spectra were traced at room temperature (RT) and at Larmor frequency of 400.13 MHz. 1H static NMR spectra were measured at 200.13 MHz in the range of 135–490 K. 1H spin-lattice relaxation times, T1, were measured at 200.13 and 19.65 MHz in the ranges of 135–490 and 153–456 K, respectively. The 1H chemical shift for the acidic proton (14.7 ppm) indicates strong hydrogen bonds. In phase III, NH4+ reorientation takes place; one type of NH4+ ions reorients with an activation energy (Ea) of 14 kJ mol 1 and the inverse of a frequency factor (τ0) of 0.85 × 10 14 s. In phase II, a very fast local and anisotropic motion of the acidic protons takes place. NH4+ ions start to diffuse translationally, and no proton exchange is observed between NH4+ ions and the acidic protons. In phase I, both NH4+ ions and the acidic protons diffuse translationally. The acidic protons diffuse with parameters of Ea = 27 kJ mol 1 and τ0 = 4.2 × 10 13 s. The translational diffusion of the acidic protons is responsible for the macroscopic proton conductivity, as the NH4+ translational diffusion is slow and proton exchange between NH4+ ions and the acidic protons is negligible.  相似文献   

19.
Colorless plate-like crystals (AlPO-1,2DAP) crystallize from a clear ‘AlPO4-1,2-diaminopropane’ reaction system. The results of elemental, thermal and FTIR analyses give the following chemical formula: (NH4+)(1,2-H2DAP)2+(AlP2O8)3−, confirmed also by single crystal structure determination. The X-ray diffraction data collected at 150 K reveal the space group Pc21n with a=8.2788(2), b=16.7882(3), c=8.6608(2) Å, and V=1203.73(5) Å3. The AlPO-1,2DAP has a chain structure related to the aluminophosphate AlPO-enA structure. Two occluded cations, NH4+ and doubly protonated diamine, balance the negative charge and interact with the chain oxygen atoms via hydrogen bond network.  相似文献   

20.
Xianghe Ren  Lihua Bai 《Optik》2012,123(11):978-981
We theoretically study the influence of the internuclear vector on molecular ionization in linear polarization laser fields through taking O2, CO2 as model molecules. We find that the ionization rates of O2 and CO2 depend on the molecular orientations. For O2, the molecular orientation corresponding to the maximum ionization rate is about φm = 45°, which is independent of the laser intensity; while for CO2, this kind of molecular orientation varies with laser intensity. We also find the ionization suppression of molecule depends on the molecular orientations and the internuclear distance. The ionization suppression easily disappears for molecules with larger internuclear distance.  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