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 15 毫秒
1.
Carbon monoxide adsorbed species on Ru/silicalite and their reactivity towards oxygen have been studied using in situ diffuse reflectance infrared spectroscopy (DRIFT). Four species were detected: a bridge bonded CO on metallic Ru (1980 cm -1), a linearly adsorbed CO on metallic Ru (2040 cm-1), a linearly adsorbed CO on partially oxidized Ru (2081 cm-1), and a multicarbonyl on oxidized Ru (2081 and 2133 cm-1). Among the four CO species the most strongly held species was the bridge bonded CO on metallic Ru while the multicarbonyl was the most weakly held species. These four species have different reactivities towards oxygen. Only the linearly adsorbed CO on metallic Ru can be oxidized at room temperature.  相似文献   

2.
The activation of adsorbed CO is an important step in CO hydrogenation. The results from TPSR of pre-adsorbed CO with H2 and syngas suggested that the presence of H2 increased the amount of CO adsorption and accelerated CO dissociation. The H2 was adsorbed first, and activated to form H* over metal sites, then reacted with carbonaceous species. The oxygen species for CO2 formation in the presence of hydrogen was mostly OH^*, which reacted with adsorbed CO subsequently via CO^*+OH^* → CO2^*+H^*; however, the direct CO dissociation was not excluded in CO hydrogenation. The dissociation of C-O bond in the presence of H2 proceeded by a concerted mechanism, which assisted the Boudourd reaction of adsorbed CO on the surface via CO^*+2H^* → CH^*+OH^*. The formation of the surface species (CH) from adsorbed CO proceeded as indicated with the participation of surface hydrogen, was favored in the initial step of the Fischer-Tropsch synthesis.  相似文献   

3.
Infrared spectra of CO-treated platinum hydrosols subsequently treated with acetylene, hydrogen, and oxygen reveal that v(CO)ads decreases from 2070 cm−1 with increasing gas-treatment time. This has been attributed to a reduction in the coverage of adsorbed CO. In Pt sol/CO/C2H2 systems, v(CO)ads decreases to a limiting value of ca. 2060 cm−1 after exposure to acetylene. In the Pt sol/CO/H2 systems, v(CO)ads decreases to ca. 2050 cm−1 after exposure to hydrogen gas. The lower frequency in the Pt sol/CO/H2 system has been attributed to CO adsorption on more active metal sites formed from the reduction of surface platinum oxides. Exposure of the CO-treated platinum hydrosols to O2 gas was found to cause the eventual disappearance of the v(CO)ads band in infrared spectra, which was attributed to oxidation of adsorbed CO to CO2 by weakly bound surface layers of platinum oxides formed by the oxygen treatment.  相似文献   

4.
Cu-based catalysts have been widely applied in electroreduction of carbon dioxide (CO2ER) to produce multicarbon (C2+) feedstocks (e.g., C2H4). However, the high energy barriers for CO2 activation on the Cu surface is a challenge for a high catalytic efficiency and product selectivity. Herein, we developed an in situ *CO generation and spillover strategy by engineering single Ni atoms on a pyridinic N-enriched carbon support with a sodalite (SOD) topology (Ni-SOD/NC) that acted as a donor to feed adjacent Cu nanoparticles (NPs) with *CO intermediate. As a result, a high C2H4 selectivity of 62.5 % and an industrial-level current density of 160 mA cm−2 at a low potential of −0.72 V were achieved. Our studies revealed that the isolated NiN3 active sites with adjacent pyridinic N species facilitated the *CO desorption and the massive *CO intermediate released from Ni-SOD/NC then overflowed to Cu NPs surface to enrich the *CO coverage for improving the selectivity of CO2ER to C2H4.  相似文献   

5.
The formation of lacunar heteropolyanions (HPA): [AsW9O33]9–, [As2W19O67(H2O)]14–, and [As2W20O68(H2O)]10– in aqueous solutions was investigated by Raman spectroscopy at [Na2HAsO3]0 = 0.1, [Na2WO4]0 = 0.9 mol L–1 and pH 9.4–1.6. The [AsW9O33]9– HPA is characterized by the most intense band ns (W=O) at 948 cm–1 retaining its position in the pH range from 8.9 to 7.5. Under these conditions, the equilibrium constant of [AsW9O33]9– formation from H2AsO3 and WO4 2– ions was estimated (logK = 87.0±1.0). The asymmetrical band at 952 cm–1 corresponding to Hx[As2W19O67(H2O)](14–x)– shifts to 960 cm–1 as the pH decreases from 6.5 to 5.5, which is due to the change in HPA protonation. The [As2W20O68(H2O)]10– HPA is formed at pH 3.1—1.6; it is characterized by a band at 972 cm–1.  相似文献   

6.
Photographic Raman spectra were obtained at shifts to ca. 7000 cm–1 for pure water and for a saturated aqueous solution of NaCl using argon ion laser excitation. Raman spectra were also obtained photoelectrically for H2O and D2O between ca. 2500 and ca. 7000 cm–1 using 248-nm excimer laser excitation and boxcar detection. Overtone and combination assignments are presented for H2O and D2O. The first IR OH-stretching overtone from water occurs 215 cm–1 above the first Raman OH-stretching overtone because the IR overtones are dominated by asymmetric stretching. The second OH-stretching Raman overtone from water is estimated to occur near 10,020 ± 20 cm–1, with 9950 cm–1 as a lower limit.  相似文献   

7.
Aggregates of CO were generated in Ar and O2 matrices by several methods, including deposition of concentrated samples, warming experiments, and photolysis of H2CO and H2C2O2. The IR absorption frequencies of the CO species were measured by an FTIR spectrometer. In solid Ar, the CO monomer was seen to absorb at 2138.5 cm?1 and (CO)2 at 2136.6 cmt-1. In systems containing only CO and Ar, no distinct peak was observed at 2140 cm?1 in contrast to previously reported results. The difference may be due to effects of H2O on CO frequencies and intensities. In solid O2, the CO monomer absorbed at 2136.6 cm?1, but it was not possible to assign a structure to (CO)2 because of a more complicated absorption spectrum than in solid Ar. The spectral data suggest that the CO dimer absorbing at 2136.6 cm?1 in solid Ar may have a side-by-side antiparallel geometric structure.  相似文献   

8.
Reacting Re(CO)5Cl with the azopyridine ligand (1) (L) in boiling benzene afford the complex Re(CO)3Cl(L), (2) in excellent yield [L=2-(p-Cl-C6H4NN)C5H4N]. The chelation of the azopyridine ligand accompanied by displacement of the two carbon monoxide ligands furnish a five-membered chelate ring. Structure determination of complex (2) has revealed a distorted octahedral ReC3N2Cl coordination sphere. The Re–N(pyridine) and, Re–N(azo) distances are 2.158(3) and 2.153(6) Å respectively, and the N–N length [1.273(4) Å], implicate relatively weak Re-azo(π*) back–bonding. The Re(CO)3Cl(L) lattice consists of C–H...Cl hydrogen bonding and Cl...O non-bonded interactions constituting a supramolecular network. Extended Hückel calculations reveal that the LUMO of Re(CO)3Cl(L) is Ca. 57% azo in character. One-electron quasireversible electrochemical reduction of the complex occurs near −0.3 V versus Saturated Calomel electrode(s.c.e.) The redox orbital is believed to belong to the above noted LUMO. Electrogenerated Re(CO)3Cl(L) underwent spontaneous solvolytic chloride displacement in MeCN, resulting in the isolation of Re(CO)3(MeCN)(L). The latter complex in turn reacted with imidazole and triphenylphosphine, furnishing Re(CO)3(C3H4N2)(L) and Re(CO)3(PPh3)(L), respectively. The pattern of carbonyl stretching frequencies of these radical anion complexes is similar to that of Re(CO)3Cl(L) but with shifts to lower frequencies by 10–20 cm−1. All three radical anion systems are one-electron paramagnetic (1.7–1.8 μB). The unpaired electron is primarily localized on the azoheterocycle ligand in a predominantly azo-π* orbital, but a small metal contribution (185, 187Re, I=5/2) is also present. Thus Re(CO)3(MeCN)(L) and Re(CO)3(C3H4N2)(L) display six-line e.p.r. spectra (A ˜ 28 G). The line shapes and intensities are characteristic of the presence of g-strain. In the case of Re(CO)3(PPh3)(L) seven nearly equispaced lines are observed due to virtually equal coupling between the metal and 31P (I=&frac;) nuclei. The g-values of the radical species are slightly higher than the free-electron value of 2.0023.  相似文献   

9.
Monooctyl--(4-carboxyanilino)benzylphosphonic acid (H2L) was investigated as a novel reagent for the extraction and separation of lead(II) and bismuth(III) from nitrate solutions. Stoichiometric ionization constants of H2L in ethanol — water mixtures determined by potentiometric NaOH titration had values pKa1 2.25 and pKa2 4.33. Ethanol and chloroform solutions of H2L followed Beer's law at 300 nm and 296 nm, molar absorptivities being 2.94·104 mol–1·l·cm–1 and 2.85·104 mol–1·l·cm–1, respectively. HNO3 and HCl were also extracted into chloroform solutions containing H2L. Bismuth(III) was quantitatively transferred into the chloroform H2L solution at pH 1, and lead(II) at 3.5. The value of the separation factor DBi/DPb is 160 at pH 1.  相似文献   

10.
Pulse radiolysis transient UV–visible absorption spectroscopy was used to study the UV–visible absorption spectrum (225–575 nm) of the phenyl radical, C6H5(), and kinetics of its reaction with NO. Phenyl radicals have a strong broad featureless absorption in the region of 225–340 nm. In the presence of NO phenyl radicals are converted into nitrosobenzene. The phenyl radical spectrum was measured relative to that of nitrosobenzene. Based upon σ(C6H5NO)270 nm=3.82×10−17 cm2 molecule−1 we derive an absorption cross-section for phenyl radicals at 250 nm, σ(C6H5())250 nm=(2.75±0.58)×10−17 cm2 molecule−1. At 295 K in 200–1000 mbar of Ar diluent k(C6H5()+NO)=(2.09±0.15)×10−11 cm3 molecule−1 s−1.  相似文献   

11.
The kinetics of the hydrogen oxidation and the CO adsorption on a Pt (ultra)microelectrode is studied in a 0.5 M H2SO4 solution saturated with a mixture of gaseous H2 and CO at partial CO pressures p CO = 10–500 ppm. The balance between rates of diffusion and adsorption of CO at different adsorption times is studied. Studied is the effect of CO impurities in H2 on steady-state polarization curves for the hydrogen ionization and nonsteady-state curves of the oxidation current decay with time at 0.02–0.05 V. Conditions under which in a certain time interval and at a certain CO concentration the slope of an I vs. t curve is proportional to p CO are determined. The obtained dependence may be used when designing a technique for monitoring CO impurities in technical hydrogen.  相似文献   

12.
The interactions of CO with a high specific surface area tin dioxide was investigated by FTIR spectroscopy and thermogravimetric analysis. FTIR study of CO interactions have shown that CO can adsorb on cus (coordinatively unsaturated sites) Sn4+ cation sites (band at 2201 cm-1). In addition, CO reacts with surface oxygen atoms. This leads to the partial reduction of SnO2 surface and to the formation of ionised oxygen vacancies together with the release of free electrons, which are responsible for the loss of transmission. Formed CO2 can chemisorb on specific surface sites: on basic sites to form carbonates species and on acidic sites (Sn4+-CO2 species) which is in competition with the formation of Sn4+-CO species. TG experiment have shown that the reduction of SnO2 by CO at 400°C occurs in two steps. First, the reduction of SnO2 surface, which is a quick phenomenon. This has allowed to evaluate that more than 12% of reducible surface oxygens can react with CO, essentially because of the presence of a large amount of surface hydroxyl groups. The second step of the reduction of SnO2 would be the progressive reduction of SnO2 bulk by the slow diffusion of oxygen atoms from the bulk to the surface. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

13.
The structures of B2H5·, B2H5CO·, and B2H5N2· radicals are investigated using the 6–31G* basis set. Both double H-bridged and single H-bridged isomers are found to be local minima on the potential energy surface. The effects of electron correlation are taken into account using single point MP4/6–31G* calculations and, for the diboryl radicals, complete MP3/6–31G* optimizations. In all cases the single H-bridged isomers are found to be more stable than the corresponding double H-bridged isomers.The transition state for the double H-bridged to single H-bridged B2H5· isomerization reaction is calculated to be 2.54 kcal mol–1 above the double H-bridged radical at the MP4SDTQ/6-31G*//UHF/ 6–31 G* level when corrected for zero point energy. Barrier tunneling increased the reaction rate by a factor of 2.5–3.0, strongly suggesting the system is fluxional at this temperature.The addition of CO and N2 to the diboryl radicals leads to relocation of the unpaired electron and rehybridization of the C and N atoms adjacent to the boron atoms. The isomers of B2H5CO· and B2H5N2· are different and should be distinguishable experimentally. While the CO moiety is bound to the diboryl radicals isomers by over 19 kcal mol–1, no binding energy is evident for N2.  相似文献   

14.
Infrared spectra of matrices codeposited Ar/HX (X=Cl, Br) with Ar/H2CCO mixtures have been examined. Isotopic substitutions (HX, DX, H2CCO, D2CCO) showed that ketene formed the 1:1 hydrogen-bonded complex with HX. The HX stretching modes were observed at 2684 cm−1 in the H2CCO–HCl complex and at 2384 cm−1 in the H2CCO–HBr complex. The ν1 modes of the ketene submolecules were shifted to low frequency and the ν9 modes to high frequency. It was proposed for the structure of the complex that the acid proton is bonded to the C=C pi electron system.  相似文献   

15.
Adsorption (at a low temperature) of nitrogen on the protonic zeolite H-Y results in hydrogen bonding of the adsorbed N2 molecules with the zeolite Si(OH)Al Brønsted-acid groups. This hydrogen-bonding interaction leads to activation, in the infrared, of the fundamental N–N stretching mode, which appears at 2334 cm−1. From infrared spectra taken over a temperature range, the standard enthalpy of formation of the OH···N2 complex was found to be ΔH0 = −15.7(±1) kJ mol−1. Similarly, variable-temperature infrared spectroscopy was used to determine the standard enthalpy change involved in formation of H-bonded CO complexes for CO adsorbed on the zeolites H-ZSM-5 and H-FER; the corresponding values of ΔH0 were found to be −29.4(±1) and −28.4(±1) kJ mol−1, respectively. The whole set of results was analysed in the context of other relevant data available in the literature.  相似文献   

16.
The variations in catalytic properties for CO hydrogenation and adsorptive properties for H2 and CO of silica-supported cobalt catalyst as a function of reduction temperature (400 to 700 °C) were investigated. A mild sintering of cobalt metal from 130 Å to 145 Å occurred only when the reduction temperature was further increased from 600 to 700 °C. However, a monotone decrease in the activity (per gram Co) for CO hydrogenation was observed with increasing reduction temperature. This is accompanied with a decrease in methane and an increase in olefin formation. The decrease in activity can not be accounted for by the variation of crystallite size of cobalt metal. A concurrent suppression in the amount of hydrogen and CO adsorptions was observed after reduction temperature was raised to 700 °C. The turnover frequencies, based on cobalt dispersion estimated from H2 and CO chemisorption, remained, more or less, similar with increasing reduction temperature from 400 to 600 °C, but increased ca. 3 and 1.5 times, respectively, at 700 °C. The relative amount of subcarbonyl species (2062 cm?1) with respect to the linear adsorbed CO (2010 cm?1) was found to increase as well. Furthermore, reduction of silica surface at 700 °C was indicated by the presence of ?Si-H, which was detected by infrared spectroscopy at 2294 cm?1. This shows the presence of metal-support interaction for silica-supported cobalt systems after reduction treatment at 700 °C.  相似文献   

17.
Cu/Cr/Ba based catalysts were found to be nonprecious metal catalysts that can selectively oxidize CO in a H2-containing stream. The CO concentration in the methanol reformer effluent can be reduced from 1–2 mol% to about 0.3 mol% with only a very small extent of H2 oxidation.  相似文献   

18.
For on-line production of [15O]–O2, [15O]–CO2, [15O]–CO, and [15O]–H2O, a new apparatus was assembled. The yield at a beam current of 10 A and a flow rate of 750 ml min–1 was about 110 MBq ml–1 target gas or 93 MBq ml–1 H2O at steady state.  相似文献   

19.
The temporal behavior of infrared spectra obtained during NO adsorption on oxidized and reduced Fe-ZSM-5 at –100, 0°C, and ambient temperature is reported. The band assignment is made based on the adsorption of labeled molecules. Bands near 1838 and 1886 cm–1 (mononitrosyl Fe2+(NO) species) form quickly and remain invariant. Bands at 1922 and 1813 cm–1 (dinitrosyl Fe2+(NO)2) together with a band at 1750 cm–1 (another mononitrosyl species) gradually become more intense for hours. Purging with He at 0–500°C leads to a gradual decrease in the intensity of all the bands. Mononitrosyl bands near 1886 and 1838 cm–1 are the most stable. The features of the IR spectra of adsorbed NO suggest the presence of dispersed Fe oxide clusters in the zeolitic pore network in addition to Fe2+ ions in cationic positions of ZSM-5.  相似文献   

20.
The specific surface and the porosity of silicate supports (SiO2, ZrO2 · SiO2, CoO · SiO2) were determined. The adsorption properties and the reducing ability of the catalysts containing 10 % Co were studied. The spectra of the thermo-programmed desorption of CO below 250°C possess two signals typical of the adsorption of the catalyst on the oxide and metal phases. The formation of liquid hydrocarbons from CO and H2 is assumed to proceed at surface bifunctional centers.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 4, pp. 668–672, April, 1993.  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