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 359 毫秒
1.
Two-dimensional graphitic metal–organic frameworks (GMOF) often display impressive electrical conductivity chiefly due to efficient through-bond in-plane charge transport, however, less efficient out-of-plane conduction across the stacked layers creates large disparity between two orthogonal conduction pathways and dampens their bulk conductivity. To address this issue and engineer higher bulk conductivity in 2D GMOFs, we have constructed via an elegant bottom-up method the first π-intercalated GMOF (iGMOF1) featuring built-in alternate π-donor/acceptor (π-D/A) stacks of CuII-coordinated electron-rich hexaaminotriphenylene (HATP) ligands and non-coordinatively intercalated π-acidic hexacyano-triphenylene (HCTP) molecules, which facilitated out-of-plane charge transport while the hexagonal Cu3(HATP)2 scaffold maintained in-plane conduction. As a result, iGMOF1 attained an order of magnitude higher bulk electrical conductivity and much smaller activation energy than Cu3(HATP)2 (σ=25 vs. 2 S m−1, Ea=36 vs. 65 meV), demostrating that simultaneous in-plane (through-bond) and out-of-plane (through πD/A stacks) charge transport can generate higher electrical conductivity in novel iGMOFs.  相似文献   

2.
Medium‐sized phosphorus cluster cations were generated by laser ablation of red phosphorus and investigated by the method of collision‐induced dissociation mass spectrometry. Experimental results show that the primary dissociation channels of phosphorus cluster cations of P + 2m+1 (6 ≤ m ≤ 11) are all characterized by the loss of P4 unit. For larger cluster cations, their dissociation pathways were more complex. For those magic cations of P + 8k+1 observed previously, their dissociation pathways progressively change from the loss of P4 unit (for k = 3) to the loss of P8 unit (for k = 4, 5). A new dissociation pathway characterized by the loss of P10 unit was also indentified for larger cations of P + 8k+1 (6 ≤ k ≤ 8). Theoretical calculation also shows that, for cations of P + 2m+1 (4 ≤ m ≤ 10), the dissociation channel characterized by the loss of P4 unit is more energetically favorable than other dissociation channels, which is in good agreement with the experimental results. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

3.
It has been demonstrated that the triplet lifetime of nonemitting molecules in the dilute vapor phase - even for complex triplet decays - can be accurately determined by means of time-resolved triplet-triplet (T-T) energy transfer to a strong emitter molecule. Besides the test molecules 1-butyne-3-one and benzaldehyde the lifetime of the vibrationally relaxed nonemitting T1(nπ*) state of cycloheptanone, τ=63 ± 5 µs at ~?0.5 Torr, together with its energy transfer rate constant to biacetyl, kET=(1.80±0.08) x 106 s?1 Torr?1, have been measured.  相似文献   

4.
The triplet-triplet energy transfer from benzaldehyde to biacetyl and the competing self-quenching between triplets and ground state molecules of benzaldehyde were investigated in the dilute vapor phase by monitoring the phosphorescence (T1(nπ*)So) decay of benzaldehyde. Following excitation into the S1(nπ*)S0 absorption band, a triplet self-quenching rate constant of kSQ=(2.4±0.1) × 104 s?1 Torr?1, corresponding to a gas-kinetic cross section of σSQ=0.22 A2, was measured. The collision-free lifetime of the benzaldehyde triplet was found to be 2.3 ± 0.4 ms. Substitution of the aldehydic proton by deuterium reduces kSQ by a factor of two: complete deuteration of the molecule has no further effect. Under the same excitation conditions, the energy transfer rate to biacetyl is kET=(2.8 ± 0.1) × 106 s?1 Torr?1, with σET = 24 A2. This process is not influenced by deuteration.  相似文献   

5.
Closely positioned donor–acceptor pairs facilitate electron‐ and energy‐transfer events, relevant to light energy conversion. Here, a triad system TPACor‐C60 , possessing a free‐base corrole as central unit that linked the energy donor triphenylamine ( TPA ) at the meso position and an electron acceptor fullerene (C60) at the β‐pyrrole position was newly synthesized, as were the component dyads TPA‐Cor and Cor‐C60 . Spectroscopic, electrochemical, and DFT studies confirmed the molecular integrity and existence of a moderate level of intramolecular interactions between the components. Steady‐state fluorescence studies showed efficient energy transfer from 1 TPA* to the corrole and subsequent electron transfer from 1corrole* to fullerene. Further studies involving femtosecond and nanosecond laser flash photolysis confirmed electron transfer to be the quenching mechanism of corrole emission, in which the electron‐transfer products, the corrole radical cation ( Cor?+ in Cor‐C60 and TPA‐Cor?+ in TPACor‐C60 ) and fullerene radical anion (C60??), could be spectrally characterized. Owing to the close proximity of the donor and acceptor entities in the dyad and triad, the rate of charge separation, kCS, was found to be about 1011 s?1, suggesting the occurrence of an ultrafast charge‐separation process. Interestingly, although an order of magnitude slower than kCS, the rate of charge recombination, kCR, was also found to be rapid (kCR≈1010 s?1), and both processes followed the solvent polarity trend DMF>benzonitrile>THF>toluene. The charge‐separated species relaxed directly to the ground state in polar solvents while in toluene, formation of 3corrole* was observed, thus implying that the energy of the charge‐separated state in a nonpolar solvent is higher than the energy of 3corrole* being about 1.52 eV. That is, ultrafast formation of a high‐energy charge‐separated state in toluene has been achieved in these closely spaced corrole–fullerene donor–acceptor conjugates.  相似文献   

6.
This paper demonstrates the effectiveness of using the redox couple to investigate DNA monolayers, and compares the potential advantages of this system to the standard redox couple. B-DNA monolayers were converted to M-DNA by incubation in buffer containing 0.4 mM Zn2+ at pH 8.6 and studied by cyclic voltammetry (CV), impedance spectroscopy (IS) and chronoamperometry (CA) with . Compared to B-DNA, M-DNA showed significant changes in CV, IS and CA spectra. However, only small changes were observed when the monolayers were incubated in Mg2+ at pH 8.6 or in Zn2+ at pH 6.0. The heterorgeneous electron-transfer rate (kET) between the redox probe and the surface of a bare gold electrode was determined to be 5.7 × 10−3 cm/s. For a B-DNA modified electrode, the kET through the monolayer was too slow to be measured. However, under M-DNA conditions, a kET of 1.5 × 10−3 cm/s was reached. As well, the percent change in resistance to charge transfer, measured by IS, was used to illustrate the dependence of M-DNA formation on pH. This result is consistent with Zn2+ ions replacing the imino protons on thymine and guanine residues. The redox couple was also effective in differentiating between single-stranded and double-stranded DNA during de-hybridization and rehybridization experiments.  相似文献   

7.
Here, the reduction chemistry of mono- and binuclear α-diimine-Re(CO)3 complexes with proton responsive ligands and their application in the electrochemically-driven CO2 reduction catalysis are presented. The work was aimed to investigate the impact of 1) two metal ions in close proximity and 2) an internal proton source on catalysis. Therefore, three different Re complexes, a binuclear one with a central phenol unit, 3 , and two mononuclear, one having a central phenol unit, 1 , and one with a methoxy unit, 2 , were utilised. All complexes are active in the CO2-to-CO conversion and CO is always the major product. The catalytic rate constant kcat for all three complexes is much higher and the overpotential is lower in DMF/water mixtures than in pure DMF (DMF=N,N-dimethylformamide). Cyclic voltammetry (CV) studies in the absence of substrate revealed that this is due to an accelerated chloride ion loss after initial reduction in DMF/water mixtures in comparison to pure DMF. Chloride ion loss is necessary for subsequent CO2 binding and this step is around ten times faster in the presence of water [ 2 : kCl(DMF)≈1.7 s−1; kCl(DMF/H2O)≈20 s−1]. The binuclear complex 3 with a proton responsive phenol unit is more active than the mononuclear complexes. In the presence of water, the observed rate constant kobs for 3 is four times higher than of 2 , in the absence of water even ten times. Thus, the two metal centres are beneficial for catalysis. Lastly, the investigation showed that the phenol unit has no impact on the rate of the catalysis, it even slows down the CO2-to-CO conversion. This is due to an unproductive, competitive side reaction: After initial reduction, 1 and 3 loose either Cl or undergo a reductive OH deprotonation forming a phenolate unit. The phenolate could bind to the metal centre blocking the sixth coordination site for CO2 activation. In DMF, O−H bond breaking and Cl ion loss have similar rate constants [ 1 : kCl(DMF)≈2 s−1, kOH≈1.5 s−1], in water/DMF Cl loss is much faster. Thus, the effect on the catalytic rate is more pronounced in DMF. However, the acidic protons lower the overpotential of the catalysis by about 150 mV.  相似文献   

8.
The gas-phase eliminations of several tert-butyl esters, in a static system and in vessels seasoned with allyl bromide, have been studied in the temperature range of 171.5–280.1°C and the pressure range of 23–98 torr. The rate coefficients for the homogeneous unimolecular elimination of these esters are given by the following Arrhenius equations: for tert-butyl pivalate, log k1(s?1) = (13.44 ± 0.30) ? (169.1 ± 3.1) kJ · mol?1 (2.303RT)?1; for tert-butyl trichloroacetate, log k1(s?1) = (12.41 ± 0.08) ? (141.1 ± 0.7) kJ · mol?1 (2.303RT)?1; and for tert-butyl cyanoacetate log k1(s?1) = (11.31 ± 0.44) ? (137.8 ± 4.1) kJ · mol?1 (2.303RT)?1. The data of this work together with those reported in the literature yield a good linear relationship when plotting log k/k0 vs. σ* values (ρ* = 0.635, correlation coefficient r = 0.972, and intercept = 0.048 at 250°C). The positive ρ* value suggests that the movement of negative charge to the acyl carbon in the transition state is rate determining. The present results along with previous investigations ratify the generalization that electron-withdrawing substituents at the acyl side of ethyl, isopropyl, and tert-butyl esters enhance the elimination rates, while electron-releasing groups tend to reduce them. The negative nature of the acyl carbon and the polarity in the transition state increases slightly from primary to tertiary esters.  相似文献   

9.
Strong push-pull interactions between electron donor, diaminoazobenzene (azo), and an electron acceptor, perylenediimide (PDI), entities in the newly synthesized A−D−A type triads (A=electron acceptor and D=electron donor) and the corresponding A−D dyads are shown to reveal wide-band absorption covering the entire visible spectrum. Electrochemical studies revealed the facile reduction of PDI and relatively easier oxidation of diaminoazobenzene in the dyads and triads. Charge transfer reversal using fluorescence-spectroelectrochemistry wherein the PDI fluorescence recovery upon one-electron oxidation, deterring the charge-transfer interactions, was possible to accomplish. The charge transfer state density difference and the frontier orbitals from the DFT calculations established the electron-deficient PDI to be an electron acceptor and diaminoazobenzene to be an electron donor resulting in energetically closely positioned PDI δ− -Azo δ+ -PDI δ− quadrupolar charge-transfer states in the case of triads and Azo δ+ -PDI δ− dipolar charge-transfer states in the case of dyads. Subsequent femtosecond transient absorption spectral studies unequivocally proved the occurrence of excited-state charge transfer in these dyads and triads in benzonitrile wherein the calculated forward charge transfer rate constants, kf, were limited to instrument response factor, meaning >1012 s−1 revealing the occurrence of ultrafast photo-events. The charge recombination rate constant, kr, was found to depend on the type of donor-acceptor conjugates, that is, it was possible to establish faster kr in the case of triads (∼1011 s−1) compared to dyads (∼1010 s−1). Modulating both ground and excited-state properties of PDI with the help of strong quadrupolar and dipolar charge transfer and witnessing ultrafast charge transfer events in the studied triads and dyads is borne out from the present study.  相似文献   

10.
The rate coefficient, k1, for the gas‐phase reaction OH + CH3CHO (acetaldehyde) → products, was measured over the temperature range 204–373 K using pulsed laser photolytic production of OH coupled with its detection via laser‐induced fluorescence. The CH3CHO concentration was measured using Fourier transform infrared spectroscopy, UV absorption at 184.9 nm and gas flow rates. The room temperature rate coefficient and Arrhenius expression obtained are k1(296 K) = (1.52 ± 0.15) × 10?11 cm3 molecule?1 s?1 and k1(T) = (5.32 ± 0.55) × 10?12 exp[(315 ± 40)/T] cm3 molecule?1 s?1. The rate coefficient for the reaction OH (ν = 1) + CH3CHO, k7(T) (where k7 is the rate coefficient for the overall removal of OH (ν = 1)), was determined over the temperature range 204–296 K and is given by k7(T) = (3.5 ± 1.4) × 10?12 exp[(500 ± 90)/T], where k7(296 K) = (1.9 ± 0.6) × 10?11 cm3 molecule?1 s?1. The quoted uncertainties are 2σ (95% confidence level). The preexponential term and the room temperature rate coefficient include estimated systematic errors. k7 is slightly larger than k1 over the range of temperatures included in this study. The results from this study were found to be in good agreement with previously reported values of k1(T) for temperatures <298 K. An expression for k1(T), suitable for use in atmospheric models, in the NASA/JPL and IUPAC format, was determined by combining the present results with previously reported values and was found to be k1(298 K) = 1.5 × 10?11 cm3 molecule?1 s?1, f(298 K) = 1.1, E/R = 340 K, and Δ E/R (or g) = 20 K over the temperature range relevant to the atmosphere. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 40: 635–646, 2008  相似文献   

11.
Two series of geometrically‐related dyads are discussed based on the difluoroborondipyrromethene (Bodipy) unit, and incorporating covalently attached hydroquinone/quinone groups. These units are anchored directly, or via a phenylene spacer, to the Bodipy core at the meso position in one series ( BD‐MHQ , BD‐MQ , BD‐MPHQ , BD‐MPQ ), but for the second series the attachment site is the 2‐position ( BD‐SHQ , BD‐SQ , BD‐SPHQ , BD‐SPQ ). The compounds show various levels of fluorescence depending on the oxidation state of the appended group and the substitution pattern. In non‐polar solvents such as toluene, diethyl ether and dichlorobenzene, the S1 state deactivation of the Bodipy unit in BD‐SPQ and BD‐MPQ is dominated by 1, 3exciplex formation, which has not been reported for Bodipy derivatives so far. In the latter molecule, the decay of the exciplex is divided between population of the Bodipy triplet state (13 %–21 %) and ground state reformation. This partitioning is not seen for the side‐on substituted derivative, BD‐SPQ , and only ground state reformation is observed following decay of the exciplex. This difference in behavior is explained by the radical‐pair inter‐system‐crossing mechanism, which more effectively operates in BD‐MPQ because of the orthogonality of the donor‐acceptor units. In the more polar solvent CH3CN all the quinone derivatives show fast formation of the charge‐separated state (kCS) followed by slower charge recombination (kCR). The ratio kCS/kCR≤80.  相似文献   

12.
Rate coefficients have been measured for the reactions of Cl atoms with methanol (k1) and acetaldehyde (k2) using both absolute (laser photolysis with resonance fluorescence) and relative rate methods at 295 ± 2 K. The measured rate coefficients were (units of 10−11 cm3 molecule−1 s−1): absolute method, k1 = (5.1 ± 0.4), k2 = (7.3 ± 0.7); relative method k1 = (5.6 ± 0.6), k2 = (8.4 ± 1.0). Based on a critical evaluation of the literature data, the following rate coefficients are recommended: k1 = (5.4 ± 0.9) × 10−11 and k2 = (7.8 ± 1.3) × 10−11 cm3 molecule−1 s−1 (95% confidence limits). The results significantly improve the confidence in the database for reactions of Cl atoms with these oxygenated organics. Rate coefficients were also measured for the reactions of Cl2 with CH2OH, k5 = (2.9 ± 0.6) × 10−11 and CH3CO, k6 = (4.3 ± 1.5) × 10−11 cm3 molecule−1 s−1, by observing the regeneration of Cl atoms in the absence of O2. Based on these results and those from a previous relative rate study, the rate coefficient for CH3CO + O2 at the high pressure limit is estimated to be (5.7 ± 1.9) × 10−12 cm3 molecule−1 s−1. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 776–784, 1999  相似文献   

13.
Absolute rate constants for the gas phase reactions of OH radicals with ethane (k1), benzene (k2), fluorobenzene (k3), chlorobenzene (k4), bromobenzene (k5), iodobenzene (k6), and hexafluorobenzene (k7) have been measured over the temperature range 234–438 K using the flash photolysis resonance fluorescence technique. The rate constants measured at room temperature (296 K), at total pressures of argon diluent between 25 and 50 Torr, were (in units of 10?13 cm3 molecule?1 s?1): k1 = (2.30 ± 0.26), k2 = (12.9 ± 1.4), k3 = (6.31 ± 0.81), k4 = (7.41 ± 0.94), k5 = (9.15 ± 0.97), k6 = (13.2 ± 1.6), and k7 = (1.61 ± 0.24), respectively. The indicated errors are our estimate of 95% confidence limits and include two standard deviations from the least-squares analysis together with an allowance for any possible systematic errors in the measurements. At elevated temperatures and under pseudo-first-order reaction conditions, non-exponential hydroxyl radical decays were observed for benzene and the monosubstituted halo-aromatics. For ethane and hexafluorobenzene, exponential decays were observed over the complete temperature range and the data were fit by the Arrhenius expressions: k1 = (8.4 ± 3.1) × 10?12 exp[(?1050 ± 100)/T] and k7 = (1.3 ± 0.3) × 10?12 exp[(?610 ± 80)/T], respectively. The results are compared with previous literature data and the mechanistic implications are discussed.  相似文献   

14.
The [2.2.2]hericene ( 6 ), a bicyclo[2.2.2]octane bearing three exocyclic s-cis-butadiene units has been prepared in eight steps from coumalic acid and maleic anhydride. The hexaene 6 adds successively three mol-equiv. of strong dienophiles such as ethylenetetracarbonitrile (TCE) and dimethyl acetylenedicarboxylate (DMAD) giving the corresponding monoadducts 17 and 20 (k1), bis-adducts 18 and 21 (k2) and tris-adducts 19 and 22 (k3), respectively. The rate constant ratio k1/k2 is small as in the case of the cycloadditions of 2,3,5,6-tetramethylidene-bicyclo [2.2.2]octane ( 3 ) giving the corresponding monoadducts 23 and 27 (k1) and bis-adducts 25 and 29 (k2) with TCE and DMAD, respectively. Constrastingly, the rate constant ratio k2/k3 is relatively large as the rate constant ratio k1/k2 of the Diels-Alder additions for 5,6,7,8-tetramethylidenebicyclo [2.2.2]oct-2-ene ( 4 ) giving the corresponding monoadducts 24 and 28 (k1) and bis-adducts 26 and 30 (k2). The following second-order rate constants (toluene, 25°) and activation parameters were obtained for the TCE additions: 3 +TCE→ 23 : k1 = 0.591±0.012 mol?1·l·s?1, ΔH=10.6±0.4 kcal/mol, and ΔS = ?24.0±1.4 cal/mol·K (e.u.); 23 +TCE→ 25 : k2=0.034±0.0010 mol?1·l·s?1, ΔH = 10.6±0.6 kcal/mol, and ΔS = ?29.7±2.0 e.u.; 4 +TCE→ 26 : k1 = 0.172±0.035 mol?1·l·s?1, ΔH 11.3±0.8 kcal/mol, and ΔS = ?24.0±2.8 e.u.; 24 +TCE→ 26 : k2 = (6.1±0.2)·10?4 mol?1·l·s?1, ΔH = 13.0±0.3 kcal/mol, and ΔS = ?29.5±0.8 e.u.; 6 +TCE→ 17 : k1 = 0.136±0.002 mol?1·l·s?1, ΔH = 11.3±0.2 kcal/mol, and ΔS = ?24.5±0.8 e.u.; 17 +TCE→ 18 : k2 = 0.0156±0.0003 mol?1·l·s?1, ΔH = 10.9±0.5 kcal/mol, and ΔS = ?30.1 ± 1.5 e.u.; 18 +TCE→ 19 : k3=(5±0.2) · 10?5 mol?1 mol?1 ·l·s?1, ΔH = 15±3 kcal/mol, and ΔS = ?28 ± 8 e.u. The following rate constants were evaluated for the DMAD additions (CD2Cl2, 30°): 6 +DMAD→ 20 : k1 = (10±1)·10?4 mol?1 · l·s?1; 20 +DMAD→ 21 : k2 = (6.5±0.1) · 10?4 mol?1 ·l·?1; 21 +DMAD→ 22 : k3 = (1.0±0.1) · 10?4 mol?1 ·l·s?1. The reactions giving the barrelene derivatives 19, 22, 26 and 30 are slower than those leading to adducts that are not barrelenes. The former are estimated less exothermic than the latter. It is proposed that the Diels-Alder reactivity of exocyclic s-cis-butadienes grafted onto bicycle [2.2.1]heptanes and bicyclo [2.2.2]octanes that are modified by remote substitution of the bicyclic skeletons can be affected by changes inthe exothermicity of the cycloadditions, in agreement with the Dimroth and Bell-Evans-Polanyi principle. Force-field calculations (MMPI 1) of 3, 4, 6 and related exocyclic s-cis-butadienes as a moiety of bicyclo [2.2.2]octane suggested single minimum energy hypersurfaces for these systems (eclipsed conformations, planar dienes). Their flexibility decreases with the degree of unsaturation of the bicyclic skeleton. The effect of an endocyclic double bond is larger than that of an exocyclic diene moiety.  相似文献   

15.
Absolute rate constants for the reactions of OH radicals with butyl ethyl ether (k1), methyl tert-butyl ether (k2), ethyl tert-butyl ether (k3) tert-amyl methyl ether (k4) and tert-butyl alcohol (k5) have been measured over the temperature range 230–372 K using a pulsed laser photolysis-laser induced fluorescence (PLP-LIF) technique. The temperature dependence of k1k5 when expressed in Arrhenius form gave: k1 = (6.59 ± 0.66) × 10 −12 exp|(362 ± 60)/T|, k2 = (5.03 ± 0.27) × 10−12 exp|&minus(133 ± 30)/T|, k3 = (4.40 ± 0.24) × 10−12 exp|(210 ± 37)/T|,k4 = (4.7 ± 0.7) × 10−12 exp|(82 ± 85)/T|, and k5 = (2.66 ± 0.48) × 10−12 exp| −(270 ± 130)/T|. However, the Arrhenius plots for k1k5, were slightly curved and are best fitted by the three parameter fits which are given in the article. The room temperature values of k1, k2, k3, k4, and k5 are (2.08 ± 0.23) × 10−11, (3.13 ± 0.36) × 10−12, (8.80 ± 0.50) × 10−12, (6.28 ± 0.45) × 10−12, and (1.08 ± 0.10) × 10−12, respectively, in cm3 molecule−1 s−1. © 1996 John Wiley & Sons, Inc.  相似文献   

16.
The rate constants for the reactions of OH with dimethyl ether (k1), diethyl ether (k2), di-n-propyl ether (k3), di-isopropyl ether (k4), and di-n-butyl ether (k5) have been measured over the temperature range 230–372 K using the pulsed laser photolysis-laser induced fluorescence (PLP-LIF) technique. The temperature dependence of k1,k4, can be expressed in the Arrhenius plots form: k1 = (6.30 ± 0.10) × 10?12 exp[?(234 ± 34)/T] and k4 = (4.13 ± 0.10) × 10?12 exp[(274 ± 26)/T]. The Arrhenius plots for k2,k3, and k5, were curved and they were fitted to the three parameter expressions: k2 = (1.02 ± 0.08) × 10?17 T2 exp[(797 ± 24)/T], k3 = (1.84 ± 0.23) × 10?17T2 exp[(767 ± 34)/T], and k5 = (6.29 ± 0.74) × 10?18T2 exp[(1164 ± 34)/T]. The values at 298 K are (2.82 ± 0.21) × 10?12, (1.36 ± 0.11) × 10?11,(2.17 ± 0.16) × 10?11, (1.02 ± 0.10) × 10?11, and (2.69 ± 0.22) × 10?11 for k1, k2, k3, k4, and k5, respectively, (in cm3 molecule?1 s?1). These results are compared to the literature data. © 1995 John Wiley & Sons, Inc.  相似文献   

17.
Relative rate coefficient data have been obtained for the reactions Br + RCHO → RCO + HBr for a series of aldehydes: HCHO, reaction (1); CH3CHO, reaction (2); CH3CH2CHO, reaction (3); CH3CH2CH2CHO, reaction (4). Measurements were made over the temperature range 240–300 K in an environmental chamber/FTIR spectrometer system, using standard relative rate techniques. All measured rate coefficient ratios were found to be independent of temperature over the range studied (k2/k1 = 3.60 ± 0.29, k3/k1 = 6.65 ± 0.53, k4/k1 = 8.62 ± 0.69, and k3/k2 = 1.80 ± 0.14), implying that the activation barriers for all four reactions are essentially identical with the A‐factors increasing with the size of the aldehyde. Relative rate coefficients for k1 and k2 agree well with currently recommended data at room temperature, but inconsistencies on the order of 20% arise at lower temperatures. The entire set of relative rate coefficient measurements is put on an absolute scale using a combination of currently recommended values for k1 and k2. The following expressions (all in units of cm3 molecule−1 s−1) are obtained: k1 = (0.79 ± 0.10) × 10−11 exp(−580 ± 200/T), k2 = (2.7 ± 0.4) × 10−11 exp(−567 ± 200/T), k3 = (5.75 ± 0.75) × 10−11 exp(−610 ± 200/T), k4 = (5.75 ± 0.75) × 10−11 exp(−540 ± 200/T), where uncertainties quoted for the A‐factor reflect the uncertainty in the room temperature value. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 460–465, 2000  相似文献   

18.
Rate constants for the reactions of OH and NO3 radicals with CH2?CHF (k1 and k4), CH2?CF2 (k2 and k5), and CHF?CF2 (k3 and k6) were determined by means of a relative rate method. The rate constants for OH radical reactions at 253–328 K were k1 = (1.20 ± 0.37) × 10?12 exp[(410 ± 90)/T], k2 = (1.51 ± 0.37) × 10?12 exp[(190 ± 70)/T], and k3 = (2.53 ± 0.60) × 10?12 exp[(340 ± 70)/T] cm3 molecule?1 s?1. The rate constants for NO3 radical reactions at 298 K were k4 = (1.78 ± 0.12) × 10?16 (CH2?CHF), k5 = (1.23 ± 0.02) × 10?16 (CH2?CF2), and k6 = (1.86 ± 0.09) × 10?16 (CHF?CF2) cm3 molecule?1 s?1. The rate constants for O3 reactions with CH2?CHF (k7), CH2?CF2 (k8), and CHF?CF2 (k9) were determined by means of an absolute rate method: k7 = (1.52 ± 0.22) × 10?15 exp[?(2280 ± 40)/T], k8 = (4.91 ± 2.30) × 10?16 exp[?(3360 ± 130)/T], and k9 = (5.70 ± 4.04) × 10?16 exp[?(2580 ± 200)/T] cm3 molecule?1 s?1 at 236–308 K. The errors reported are ±2 standard deviations and represent precision only. The tropospheric lifetimes of CH2?CHF, CH2?CF2, and CHF?CF2 with respect to reaction with OH radicals, NO3 radicals, and O3 were calculated to be 2.3, 4.4, and 1.6 days, respectively. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 42: 619–628, 2010  相似文献   

19.
An assembly has been synthesised that consists of four units: a meso‐substituted corrole (C3), perylene bisimide (PI), and two electron‐rich triphenylamine (DPA) units. PI is connected through a 1,4‐phenylene bridge to C3, whereas the two DPA units are linked to C3 through a diphenyl ether linkage, which is used for the first time to connect the various moieties. Various synthetic strategies were elaborated, and the chosen one afforded the final system in six steps in an overall yield of 6 %. The resulting assembly, made of three different units, was named a “triad”. Excitation of the corrole (C3) or perylene bisimide (PI) units led to the charge‐separated state DPA‐C3+‐PI? with a rate k>1011 s?1 in benzonitrile and dichloromethane (CH2Cl2) or with k of the order of 1010 s?1 in toluene. The latter charge‐separated state decayed to the ground state with a rate k=1.8×109 s?1 in toluene. In the polar solvents benzonitrile and dichloromethane, recombination to the ground state competes with a charge shift to form the distal charge‐separated state, DPA+‐C3‐PI?, the formation of which occurs with a yield of 50 %. Recombination to the ground state of DPA+‐C3‐PI? occurs with a rate k=5×107 s?1 in CH2Cl2 and k=2×107 s?1 in benzonitrile.  相似文献   

20.
The rate coefficients for the reactions of OH with ethane (k1), propane (k2), n-butane (k3), iso-butane (k4), and n-pentane (k5) have been measured over the temperature range 212–380 K using the pulsed photolysis-laser induced fluorescence (PP-LIF) technique. The 298 K values are (2.43±0.20) × 10?13, (1.11 ± 0.08) × 10?12, (2.46 ± 0.15) × 10?12, (2.06 ± 0.14) × 10?12, and (4.10 ± 0.26) × 10?12 cm3 molecule?1 s?1 for k1, k2, k3, k4, and k5, respectively. The temperature dependence of k1 and k2 can be expressed in the Arrhenius form: k1 = (1.03 ± 0.07) × 10?11 exp[?(1110 ± 40)/T] and k2 = (1.01 ± 0.08) × 10?11 exp[?(660 ± 50)/T]. The Arrhenius plots for k3k5 were clearly curved and they were fit to three parameter expressions: k3 = (2.04 ± 0.05) × 10?17 T2 exp[(85 ± 10)/T] k4 = (9.32 ± 0.26) × 10?18 T2 exp[(275 ± 20)/T]; and k5 = (3.13 ± 0.25) × 10?17 T2 exp[(115 ± 30)/T]. The units of all rate constants are cm3 molecule?1 s?1 and the quoted uncertainties are at the 95% confidence level and include estimated systematic errors. The present measurements are in excellent agreement with previous studies and the best values for atmospheric calculations are recommended. © 1994 John Wiley & Sons, Inc.  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