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 726 毫秒
1.
The thermal dissociation of gaseous Mo(CO)6 and W(CO)6 in an argon carrier gas, Mo(CO)6 → Mo(CO)5 + CO (1) and W(CO)6 → W(CO)5 + CO (2), is studied over temperature ranges of ∼585–685 K for (1) and ∼690−810 K for (2) at a total gas concentrations of 4 × 10−6 and 4 × 10−5 mol/cm3 by using the shock tube technique in conjunction with absorption spectrophotometry. The measured rate constants are extrapolated to the high-pressure limit by means of a newly developed procedure, with the resultant expressions for the indicated temperature ranges reading as kd1,∞(T),[s−1] = 1016.12 ± 0.68exp[(−148.8 ± 8.1 kJ/mol)/RT] and kd2,∞(T),[s−1] = 1015.93 ± 0.63exp[(−171.7 ± 8.9 kJ/mol)/RT]. Comparison of the high-pressure dissociation rate constants with the published data revealed a considerable discrepancy, a tentative explanation of which is given. Based on the obtained high-pressure dissociation rate constants and the available data on the high-pressure room-temperature rate constants for the reverse reaction of recombination, the first bond dissociation energies for these molecules are evaluated and compared with previous determinations, both theoretical and experimental. The enthalpies of formation of Mo(CO)5 and W(CO)5 are determined: ΔfH°(Mo(CO)5, g, 298.15 K) = −644.1 ± 5.6 kJ/mol and ΔfH°(W(CO)5, g, 298.15 K) = −581.9 ± 6.6 kJ/mol. Based on the enthalpies of formation of Mo(CO)5, W(CO)5, Mo(CO)6, and W(CO)6, and the published molecular parameters of these four species, their thermochemical functions are calculated and presented in the form of NASA seven-term polynomials.  相似文献   

2.
The rate constants for the reaction of CN with N2O and CO2 have been measured by the laser dissociation/laser-induced fluorescence (two-laser pump-probe) technique at temperatures between 300 and 740 K. The rate of CN + N2O was measurable above 500 K, with a least-squares averaged rate constant, k = 10−11.8±0.4 exp(−3560 ± 181/T) cm3/s. The rate of CN + CO2, however, was not measurable even at the highest temperature reached in the present work, 743 K, with [CO2] ⩽ 1.9 × 1018 molecules/cm3. In order to rationalize the observed kinetics, quantum mechanical calculations based on the BAC-MP4 method were performed. The results of these calculations reveal that the CN + N2O reaction takes place via a stable adduct NCNNO with a small barrier of 1.1 kcal/mol. The adduct, which is more stable than the reactants by 13 kcal/mol, decomposes into the NCN + NO products with an activation energy of 20.0 kcal/mol. This latter process is thus the rate-controlling step in the CN + N2O reaction. The CN + CO2 reaction, on the other hand, occurs with a large barrier of 27.4 kcal/mol, producing an unstable adduct NCOCO which fragments into NCO + CO with a small barrier of 4.5 kcal/mol. The large overall activation energy for this process explains the negligibly low reactivity of the CN radical toward CO2 below 1000 K. Least-squares analyses of the computed rate constants for these two CN reactions, which fit well with experimental data, give rise to for the temperature range 300–3000 K.  相似文献   

3.
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   

4.

The synthesis and structural characterization of a novel In(III) complex is described. The reaction between InCl3 with sodium mercapto-acetic acid (NaSCH2(CO)OH) in 4-methylpyridine (CH3(C5H5N), (4-Mepy)) at 25°C affords [ClIn(SCH2(CO)O)2]2-[(4-MepyH)2]2+ (1). X-ray diffraction studies of (1) show it to have a distorted square-pyramidal geometry with the [(-SCH2(CO)CO-)] ligands in a trans conformation. The compound crystallizes in the P1(No. 2) space group with a = 7.8624(6) Å, b = 9.950(1) Å, c = 13.793(2) Å, α = 107.60(1)°, β = 90.336(8)°, γ = 98.983(9)°, V = 1014.3(4) Å3, R(F°) = 0.037 and Rw = 0.048.  相似文献   

5.
The chemistry of the HC(O)CO radical, produced in the oxidation of glyoxal, has been studied under conditions relevant to the lower atmosphere using an environmental chamber/Fourier Transform infrared spectrometric system. The chemistry of HC(O)CO was studied over the range 224–317 K at 700 Torr total pressure and was found to be governed by competition between unimolecular decomposition [to HCO and CO, reaction (5)] and reaction with O2 [to form HO2 and 2CO, reaction (6a), or HC(O)C(O)O2, reaction (6b)]. The rate coefficient for decomposition relative to that of reaction with O2 increases with increasing temperature. Assuming a value for k6 of 10−11 cm3 molecule−1 s−1, the following expression for the unimolecular decomposition is obtained at 700 Torr, k5 = 1.4+9/−1.1 × 1012 exp(−3160 ± 500/T). The rate coefficients for reactions (6a) and (6b) are about equal, with no strong dependence on temperature. The reaction of HC(O)C(O)O2 with NO2 was also studied. Final product analysis was consistent with the formation of HCO, CO2, and NO3 as the major products in this reaction; no evidence for the PAN‐type species, HC(O)C(O)O2NO2, was found even at the lowest temperature studied (224 K). The UV‐visible absorption spectrum of glyoxal is also reported; results are in substantive agreement with previous studies. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 149–156, 2001  相似文献   

6.
The rate constants for proton transfer from H3+ ions to N2, O2, and CO have been measured as function of hydrogen buffer gas partial pressure. The rate constant for proton transfer from H3+ to N2 shows a very large pressure dependence, increasing from 1.0 × 10?9 cm3/s at low H2 partial pressures to 1.7 × 10?9 cm3/s at high H2 partial pressures. The rate constants for proton transfer from H3+ to O2 and CO are constant with partial pressure of H2; giving values of 6.4 × 10?10 cm3/s and 1.7 × 10?9 cm3/s, respectively. The roles of excess vibrational energy in H3+ ions and of equilibrium between forward and back reaction are discussed. Back reaction is observed only for the reaction of H3+ ions with O2, and an equilibrium constant of K = 2.0 ± 0.4 at 298 K has been determined. From these data the proton affinity of O2 is deduced to be 0.47 ± 0.11 kcal/mole higher than that of H2.  相似文献   

7.
Long-path FTIR spectroscopy was used to study the kinetics and mechanism of the reaction of Cl atoms with CO in air. The relative rate constants at 298 K and 760 torr for the forward direction of the reaction of Cl with 13CO and the reaction of Cl13CO with O2 were k1 = (3.4 ± 0.8) × 10−14 cm3 molecule−1 s−1 and k2 = (4.3 ± 3.2) × 10−13 cm3 molecule−1 s−1, respectively (all uncertainty limits are 2σ). The rate constant for the net loss of 13CO due to reaction with Cl in 1 atm of air at 298 K was kCl+COobs = (3.0 ± 0.6) × 10−14 cm3 molecule−1 s−1. The only observed carbon-containing product of the Cl + 12CO reaction was 12CO2, with a yield of 109 ± 18%. Our results are in good agreement with extrapolations from previous studies. The reaction mechanism and the implications for laboratory studies and tropospheric chemistry are discussed. © 1996 John Wiley & Sons, Inc.  相似文献   

8.
Heterogeneous recombination of O + CO → CO2 over a solid CO2 surface at 77 K was investigated. A modified discharge flow setup was used to generate low O atom concentrations by the reaction N + NO → N2 + O(3P). The O atom concentrations were measured upstream and downstream of the solid CO2 substrate using resonance fluorescence by monitoring the unresolved 130.3 nm triplet transition 3S1 ? 3P2,1,0 at the two fixed points. CO2 formed was determined by measuring the β activity from C14O2 produced from CO containing C14O as a reactant gas. The CO2 formation was found to be first order in CO and independent of O atom concentration over the entire range of 4.3 × 1012 to 1.9 × 1014 cm?3 and 1.2 × 1011 to 5.6 × 1012 cm?3 for CO and O respectively. The first order recombination coefficient, λCO was found to be 1.4 (±.38) × 10?5.  相似文献   

9.
Summary Fac-[188Re(CO)3(H2O)3]+ was synthesized with an overall radiochemical yield of 80±5%, and more than 95% radiochemical purity after a QMA Sep-Pak column separation. Fac-[Re(CO)3(H2O)3]+ was also synthesized as a reference sample. The structure of the precursor, fac-[188Re(CO)3(H2O)3]+, was confirmed by high performance of liquid chromatography (HPLC). MN-His (magnetic nanoparticles coated with silica and modified with an amino silane coupling agent, N-[3-(trimethyoxysilyl)propyl]-ethylenediamine (SG-Si900) and immobilized with histidine) was labeled with fac-[188Re(CO)3(H2O)3]+ and an initial animal test of MN-His was conducted for a magnetic targeting study.  相似文献   

10.
Electronically excited carbon dioxide (CO2*) is known for its broadband emission, and its detection can lead to valuable information; however, owing to its broadband characteristics, CO2* is difficult to isolate experimentally, and its chemical kinetics are not well known. Although numerous works have monitored CO2* chemiluminescence, a full kinetic scheme for the excited species has yet to be developed. To this end, a series of shock‐tube experiments was performed in H2‐N2O‐CO mixtures highly diluted in argon at conditions where emission from CO2* could be isolated and monitored. These results were used to evaluate the kinetics of CO2*, in particular the main CO2* formation reaction CO + O + M CO2* + M (R1). Based on collision theory, the quenching chemistry of CO2* was estimated for 11 collision partners. The final mechanism developed for CO2* consists of 14 reactions and 13 species. The rate for (R1) was determined to within about ±60% using low‐pressure experiments performed in five different (H2‐)N2O‐CO‐Ar mixtures, as follows: where R is the universal gas constant in cal/mol‐K and T is the temperature in K. Final mechanism predictions were compared with experiments at low and high pressures, with good agreement at both conditions for the temperature dependence of the peak CO2* and the CO2* species time histories. Comparisons were also made with previous experiments in methane–oxygen mixtures, where there was slight overprediction of CO2* experimental trends, but with the results otherwise showing a dramatic improvement over an earlier mechanism. Experimental results and model predictions were also compared with past literature rates for CO2*, with good agreement for peak CO2* trends and slight discrepancies in CO2* species time histories. Overall, the ability of the CO2* mechanism developed in this work to reproduce a range of experimental trends represents an important improvement over the existing knowledge base on chemiluminescence chemistry.  相似文献   

11.

The synthesis and structural characterization of a novel ionic Ga(III) five-coordinate complex [{CH3(C5H4N)}Ga(SCH2(CO)O)2]?[(4-MepyH)]+, (4-Mepy=CH3(C5H5N)) from the reaction between Ga2Cl4 with sodium mercapto-acetic acid in 4-methylpyridine is described. Under basic reaction conditions the mercapto ligand is found to behave as a 2e? bidentate ligand. Single crystal X-ray diffraction studies show the complex to have a distorted square-pyramidal geometry with the [(?SCH2(CO)CO?)] ligands trans. The compound crystallizes in the P21/c (No. 14) space group with a=7.7413(6)Å, b=16.744(2)Å, c=14.459(2)Å, V,=1987.1(6)&Aringsup3;, R(F, o)=0.032 and RW =0.038.  相似文献   

12.
The reaction of rhodium(I) carbonyl chloride, [Rh(CO)2Cl]2, with dichromate, cerium(IV) sulfate, hexachloroplatinic acid or p-benzoquinone in aqueous hydrochloric acid proceeds by consumption of 4 equivalents of oxidizing agent per mole or rhodium(I) in accordance with the equation RhI(CO)2  4e + H2O → RhIII(CO) + 2H+ + CO2A “cyclic” oxidation mechanism is suggested.  相似文献   

13.
A detailed theoretical survey of the potential energy surface (PES) for the CH2CO + O(3P) reaction is carried out at the QCISD(T)/6‐311+G(3df,2p)//B3LYP/6‐311+G(d,p) level. The geometries, vibrational frequencies, and energies of all stationary points involved in the reaction are calculated at the B3LYP/6‐311+G(d,p) level. More accurate energy information is provided by single‐point calculations at the QCISD(T)/6‐311+G(3df,2p) level. Relationships of the reactants, transition states, intermediates, and products are confirmed by the intrinsic reaction coordinate (IRC) calculations. The results suggest that P1(CH2+CO2) is the most important product. This study presents highlights of the mechanism of the title reaction. © 2005 Wiley Periodicals, Inc. Int J Quantum Chem, 2005  相似文献   

14.
N2O was photolyzed at 2139 Å to produce O(1D) atoms in the presence of H2O and CO. The O(1D) atoms react with H2O to produce HO radicals, as measured by CO2 production from the reaction of OH with CO. The relative importance of the various possible O(1D )–H2O reactions is The relative rate constant for O(1D) removal by H2O compared to that by N2O is 2.1, in good agreement with that found earlier in our laboratory. In the presence Of C3H6, the OH can be removed by reaction with either CO or C3H6: From the CO2 yield, k3/k2 = 75,0 at 100°C and 55.0 at 200°C to within ± 10%. When these values are combined with the value of k2 = 7.0 × 10?13exp (–1100/RT) cm3/sec, k3 = 1.36 × 10?11 exp (–100/RT) cm3/sec. At 25°C, k3 extrapolates to 1.1 × 10?11 cm3/sec.  相似文献   

15.
The reaction of N2O with CO, catalyzed by Fe+(C6H6) and producing N2 and CO2, has been investigated at the UB3LYP/6-311+G(d) level. The computation results revealed that the reaction of Fe+(C6H6), N2O and CO, is an O-atom abstraction mechanism. For the reaction channels, the geometries and the vibrational frequencies of all species have been calculated and the frequency modes analysis also have been given to elucidate the reaction mechanism. On the basis for geometry optimizations, the thermodynamic data of these reactions channels have been calculated using the statistical theory at 295.15 K and pressure of 0.35 Torr. Using Eyring transition state theory with Wigner correction, the activation thermodynamic data, rate constant and frequency factors for the these reaction channels also have been given. The results showed that CO and N2O do not react without catalyst and Fe+(C6H6) can excellently mediate the reaction of N2O and CO.  相似文献   

16.
The blue copper complex [Cu2(H2O)2(phen)2(OH)2][Cu2(phen)2(OH)2(CO3)2] · 10 H2O, which was prepared by reaction of 1,10‐phenanthroline monohydrate, CuCl2 · 2 H2O and Na2CO3 in the presence of succinic acid in CH3OH/H2O at pH = 13.0, crystallized in the triclinic space group P1 (no. 2) with cell dimensions: a = 9.515(1) Å, b = 12.039(1) Å, c = 12.412(2) Å, α = 70.16(1)°, β = 85.45(1)°, γ = 81.85(1)°, V = 1323.2(2) Å3, Z = 1. The crystal structure consists of dinuclear [Cu2(H2O)2(phen)2(OH)2]2+ complex cations, dinuclear [Cu2(phen)2(OH)2(CO3)2]2– complex anions and hydrogen bonded H2O molecules. In both the centrosymmetric dinuclear cation and anion, the Cu atoms are coordinated by two N atoms of one phen ligand, three O atoms of two μ‐OH groups and respectively one H2O molecule or one CO32– anion to complete distorted [CuN2O3] square‐pyramids with the H2O molecule or the CO32– anion at the apical position (equatorial d(Cu–O) = 1.939–1.961 Å, d(Cu–N) = 2.026–2.051 Å and axial d(Cu–O) = 2.194, 2.252 Å). Two adjacent [CuN2O3] square pyramids are condensed via two μ‐OH groups. Through the interionic hydrogen bonds, the dinuclear cations and anions are linked into 1D chains with parallel phen ligands on both sides. Interdigitation of phen ligands of neighboring 1D chains generated 2D layers, between which the hydrogen bonded water molecules are sandwiched.  相似文献   

17.
Enthalpies of solution of CO2(g), NaHCO3(s), and Na2CO3(s) in excess NaOH solution were measured at 298.15 K by solution calorimetry. The results were reduced to standard-state processes through use of results from a preceding paper, and standard enthalpies of solution for CO2(g), NaHCO3(s), and Na2CO3(s) in water at 298.15 K were found to be: ?(4720 ± 40), (4474 ± 30), and ?(6371 ± 30) calth mol?1 respectively. The results of equilibrium studies involving CO2(g) (solubility and e.m.f. studies) were reviewed and assembled,together with entropies for related solids. Standard values of ΔHfo, ΔGfo, and So at 298.15 K were evaluated for CO2(aq, non-ionized), HCO3?(aq), CO32?(aq), NaHCO3(s), Na2CO3(s), Na2CO3·H2O(s), and Na2CO3·10H2O(s).  相似文献   

18.

Ligand substitution of trans-[CoIII(en)2(Me)H2O]2+ was studied for pyrazole, 1,2,4-triazole and N-acetylimidazole as entering nucleophiles. These displace the coordinated H2O molecule trans to the methyl group to form trans-[Co(en)2(Me)azole]. Stability constants at 18°C for the substitution of H2O by pyrazole, 1,2,4-triazole and N-acetylimidazole are 0.7 ± 0.1, 13.8 ± 1.4 and 1.7 ± 0.2 M?1, respectively. Second order rate constants at the same temperature for the reaction of trans-[CoIII(en)2(Me)H2O]2+ with pyrazole, 1,2,4-triazole and N-acetylimidazole are 161 ± 12, 212 ± 11 and 12.9 ± 1.6 M?1 s?1, respectively. Activation parameters (ΔH, ΔS) are 67 ± 6 kJ mol?1, + 27 ± 19 J K?1 mol?1; 59 ± 2 kJ mol?1, + 1 ± 6 J K?1 mol?1 and 72 ± 4 kJ mol?1, + 23 ± 14 J K?1 mol?1 for reactions with pyrazole, 1,2,4-triazole and N-acetylimidazole, respectively. Substitution of coordinated H2O by azoles follows an Id mechanism.  相似文献   

19.
Counterpoise corrected ab initio calculations are reported for (H2O)2 and H2O-H2CO. Geometry searches were done in the moment-optimized basis DZP' at the SCF, MP2, and CEPA-1 levels of theory, followed by more accurate single-point calculations in basis ESPB, which includes bondfunctions to saturate the dispersion energy. The final equilibrium binding energies obtained are ?4.7 ±0.3 kcal/mol for a near-linear (H2O)2 structure and ?4.6 ±0.3 kcal/mol for a strongly bent HOH ‥ OCH2 structure. The energy difference between these systems is much smaller than in all previous ab initio work. Cyclic (C2h) and bifurcated (C2v) transition structures for (H2O)2 are located at 1.0 ±0.1 kcal/mol and 1.9 ±0.3 kcal/mol above the global minimum, respectively. A new partitioning scheme is presented that rigorously partitions the MP2 correlation interaction energy in intra and intermolecular (dispersion) contributions. These terms are large (up to 2 kcal/mol) but of opposite sign for most geometries studied and hence their overall effect upon the final structures is relatively small. The relative merits of the MP2 and CEPA-1 approaches are discussed are discussed and it is concluded that for economical reasons MP2 is to be preferred, especially for larger systems.  相似文献   

20.
Ab initio calculations using the unscaled 4-31G basis set have been carried out on the cc, tc, and tt conformers of carbonic acid and the bicarbonate ion, with full geometry optimization assuming the structures to be planar. The complete harmonic force field is reported for the (most stable) tt conformer and for the bicarbonate ion, also selected quadratic force constants for the cc and tc conformers. The changes in certain bond lengths and stretching force constants in the cctc, tctt, and cctt conformer conversion reactions are indicative of intramolecular hydrogen bonding, C?O…H? O and H? O…H? O, which is examined in greater detail by partitioning the overall conformer conversion energy into distortion and bonding energy components. The fundamental vibration frequencies for the tt conformer and the bicarbonate ion are calculated from the force constant matrices, and hence, using a scaling factor based on a comparison of calculated and experimental values for the bicarbonate ion and trans-formic acid, a value is predicted for the zero-point energy of the tt conformer. A new estimate of ΔH? for the hydration reaction, H2O + CO2 → H2CO3, at 298 K in the gas phase; is made from thermochemical data, +20.2 ± 3.4 kJ mol?1, which, together with estimates of (H298? – H0?) and the zero-point energy for H2CO3, gives +8.1 ± 7.0 kJ mol?1 for ΔET(expt). ΔET calculated from the 4-31G basis set data is -29.1 kJ mol?1. Comparison of the experimental value, the Hartree–Fock limit value, and values calculated with a variety of basis sets for the bond separation reaction, CO2 + CH4 → 2H2CO, suggests that the differences, ΔET(expt) minus ΔET(SCF ), are due mainly to basis set limitations and not substantial correlation energy contributions.  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