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.
The reaction of [Ni(COD)2] with one equivalent of DABMes (DABMes = (2,4,6‐Me3C6H2)N=C(Me)‐C(Me)=N(2,4,6‐Me3C6H2)) affords a mixture of the compound [Ni(DABMes)2] ( 2 ) and starting material [Ni(COD)2]. The crystallographically characterized, diamagnetic complex 2 can be obtained in a stoichiometric reaction of [Ni(COD)2] and two equivalents of DABMes. This reaction can be accelerated by addition of 1‐chloro‐fluorobenzene or methyl iodide. In the presence of 1‐chloro‐fluorobenzene, [Ni(DABMes)(COD)] ( 3 ) is available via reaction of [Ni(COD)2] and one equivalent of DABMes. The crystallographically characterized complex 3 reacts with diphenylacetylene to afford [Ni(DABMes)(Ph‐C≡C‐Ph)] ( 4 ). A long‐wavelength absorption band in the UV‐Vis spectrum of this compound has to be assigned to a mixed MLCT/LL′CT transition, as quantum chemical calculations reveal.  相似文献   

2.
The mechanism of the carbonylation reaction of allyl halides catalyzed by nickel (Ni(CO)4) and palladium (Cl2Pd(PPh3)2) complexes is theoretically investigated at the DFT level using the hybrid B3LYP functional. The favored reaction path to carbonylation corresponds, for both catalysts, to a direct attack of the halogen on the metal. This affords η1 intermediates that can undergo the final carbonylation step. It is also possible to obtain the acyl product (β,γ-unsaturated acyl halides) from η2 and/or η3 intermediates. However, in this case, the barrier of the rate-determining step to carbonylation is much higher. Since a channel on the potential surface connects rather easily the η2 or η3 intermediates to the η1 intermediates, an alternative and competitive path leading to the acyl products can originate from the η2 or η3 intermediates, follow the η23 → η1 transformation, then undergo the final carbonylation step.  相似文献   

3.
Ab initio calculations at the B3LYP/6–311 ++G(2df,2p) and B3LYP/6–31G(d) level have been carried out to investigate the reaction mechanism of methionine sulfoxide reductases of class A. These enzymes reduce oxidized methionine in vivo and therefore play an important role in repairing protein damage caused by the oxidative stress. Our calculations have been carried out for a model reaction in a model active site. Several reaction mechanisms have been explored that can roughly be described as (2H+ + 2e) or (H+ + e). The results suggest that the actual reaction mechanism is of the (2H+ + 2e) type corresponding to a more or less asynchronous-concerted double-proton transfer reaction leading to the formation of methionine (dimethylthioether in our model) and a sulfenic acid Cys-SOH. The Michaelis complex would involve one deprotonated Cys and one protonated Glu residues in the active site, this protonation state being mandatory to stabilize the sulfoxide substrate. Then, proton transfer from Glu to the substrate takes place, followed by proton transfer from one Tyr residue and fast reorganization of the system. The overall activation energy barrier is estimated to fall in the range 7–9 kcal/mol, much lower than the predicted barrier in DMSO solution (29.6 kcal/mol) reported before.  相似文献   

4.
《中国化学快报》2023,34(5):107752
Owing to its outstanding photoactivity, ferrioxalate is originally used as an actinometer and subsequent work has discovered that photochemistry of ferrioxalate is also fundamentally or technically important in atmospheric chemistry and water treatment. While the overall products generated from photolysis of ferrioxalate are known to include Fe(II), a series of oxidizing (e.g., ?OH, O2??/HO2??) or reducing (C2O4??/CO2??) radicals and H2O2, however, at the molecular level, the primary step of the photoreaction of ferrioxalate remains as an unsolved mystery due to the difficulty in examining such ultrafast processes. Benefiting from the development of time-resolved spectroscopy, this old question has been studied with increasing vigor recently, by means of such ever-more-sophisticated techniques (e.g., flash photolysis, time-resolved X-ray absorption spectroscopy (XAS), femtosecond infrared (IR) absorption spectroscopy, ultrafast photoelectron spectroscopy (PES)). There are two contrary views on the primary reaction mechanism: (1) Intramolecular electron transfer (ET) precedes the cleavage of the metal-ligand bond; (2) The dissociation of C–C or Fe–O bond occurs before intramolecular ET. Thus, this review presents a comprehensive summary about the overall reaction mechanism and molecular level mechanism of ferrioxalates. In chronological order, we have elaborated two predominant but controversial views from the perspectives of different experimental approaches. Some challenges and research opportunities in this active field are also briefly discussed.  相似文献   

5.
With Li+ or TBA+ as counter ion the electrodimerization of phenylpent-2-en-3-one occurs by a 0.2 F initiation. This reaction is regio- and stereospecific with Li+, the radical autocatalytic reaction gave 100% yield of a new dimer. If TBA+ is used the primary step is an ion-substrate reaction giving two isomers in equal proportion. All reaction products were structurally determined by 2D-NMR (SECSY or COSY and NOESY sequence).  相似文献   

6.
Non‐heme high‐spin (hs) {FeNO}8 complexes have been proposed as important intermediates towards N2O formation in flavodiiron NO reductases (FNORs). Many hs‐{FeNO}8 complexes disproportionate by forming dinitrosyl iron complexes (DNICs), but the mechanism of this reaction is not understood. While investigating this process, we isolated a new type of non‐heme iron nitrosyl complex that is stabilized by an unexpected spin‐state change. Upon reduction of the hs‐{FeNO}7 complex, [Fe(TPA)(NO)(OTf)](OTf) ( 1 ), the N‐O stretching band vanishes, but no sign of DNIC or N2O formation is observed. Instead, the dimer, [Fe2(TPA)2(NO)2](OTf)2 ( 2 ) could be isolated and structurally characterized. We propose that 2 is formed from dimerization of the hs‐{FeNO}8 intermediate, followed by a spin state change of the iron centers to low‐spin (ls), and speculate that 2 models intermediates in hs‐{FeNO}8 complexes that precede the disproportionation reaction.  相似文献   

7.
The concept of electron transfer catalysis (ETC), or more specifically “Double Activation Induced by Single Electron Transfer” (DAISET) gives an opportunity to connect experimental facts never previously correlated. The first activation results from the transfer of an electron to (or from) a molecular species; the second activation results from the build-up of a reaction chain able to reproduce the species formed in the first step. The starting point of this review is the SRN 1 mechanism where principle and experimental diagnostic criteria are critically discussed. The thermal and photochemical exchange and substitution reactions of PtIV complexes are then reviewed together with the exchange reaction [AuCl4]?/Cl?, reactions with Grignard reagents and other organometallic reagents, as well as the redox behavior of electronically excited organic compounds. Photochemical applications, including solar energy conversion are discussed. New aspects are also presented for the mechanistic problem “SN 2 reaction or SET process?” Moreover, the concept has significance for SH2 reactions at metal centers, molecule-induced homolyses, reactions of complexes, as well as electrochemical processes.–Unless otherwise specified, only double activation (DAISET) processes will be discussed in this article.  相似文献   

8.
Reactions between CF2 and O(3P) have been studied at 295 K in a gas flow reactor sampled by a mass spectrometer. The major reaction for CF2 has been found to be $$CF_2 + O \to COF + F$$ with $$CF_2 + O \to CO + 2F(F_2 )$$ more than a factor of three slower. The rate coefficient for all loss processes for CF2 on reaction with O is (1.8±0.4)×10?11 cm3 s?1. The COF produced in (18) undergoes a fast reaction with O to produce predominantly CO2. $$COF + O \to CO_2 + F$$ It is uncertain from the results whether or not $$COF + O \to CO + FO$$ occurs, but in any event (19) is the major route. The rate coefficient for the loss of COF in this system [i.e., the combined rate coefficients for (19) and (20)] is (9.3±2.1)×10?11 cm3 s?1. Stable product analysis reveals that for each CF2 radical consumed, the following distribution of stable products is obtained: COF2 (0.04±0.02), CO (0.21±0.04), and CO2 (0.75±0.05). Thus COF2, which we assume is produced via $$CF_2 + O \xrightarrow{M} COF_2$$ is a very minor product in this reaction sequence. The measured rate coefficients demonstrate that reactions (18) and (19) are important sources of F atoms in CF4/O2 plasmas.  相似文献   

9.
A practical procedure for the determination of branching ratios for reactions which lead to either excited or electronically ground state products is outlined. The method is applied to four reactions which could (on energetic grounds) produce an electronically excited iodine atom. No case of a complete inversion is found, but one reaction (F + HI) is predicted to yield a statistical, (one half), I*(2P1/2) to I(2P3/2) ratio.  相似文献   

10.
Hitherto there was no reaction known that permits transformations of R1R2-CO → 0.5 R1R2R3C–CR1R2R3 in one step. This type of additive–reductive carbonyl dimerization is now possible using alkoxy(alkyl)tungsten(v) complexes with aromatic, heteroaromatic or α,β-unsaturated aldehydes and ketones. When a corresponding phenyl complex was employed in a test experiment, it was revealed that an aliphatic ketone could be used as the substrate in this reaction. A second interesting type of reaction is the transformation of CH3 ligands into μ-CH2 ligands, which occurs during the treatment of MeLi or Me3Al with molybdenum or tungsten chlorides (oxidation states VI and V, for Mo additionally IV) at low temperatures with liberation of CH4. Here, the question arises as to whether the intermediate involved has a terminal CH2 ligand (Schrock carbene complex) or a μ-CH3 ligand (CH3 bound by a two-electron three-center bond to two metal atoms). Of all the μ-CH2 complexes obtained, those which were synthesized by the action of MeLi on molybdenum chlorides can be recommended as reagents for carbonylmethylenation of aldehydes and ketones. They display high selectivity, very low basicity, a surprising resistance to protons, they are readily available, can be easily modified and, as regards their selective behavior, they have been investigated more thoroughly than other readily accessible carbonylmethylenation reagents of comparable selectivity. The results of NMR spectroscopic investigations on the structure of the μ-CH2 complexes, and associated reaction mechanisms are discussed. A survey of carbonylmethylenation reagents, which have been reported in the literature, permits comparisons to be made with carbonylmethylenating molybdenum and tungsten complexes.  相似文献   

11.
Chemiluminescence is observed in the thermal reaction of phenylchlorocarbene or (2-chlorophenyl)carbene and O2, matrix-isolated in Ar. The chemiluminescence spectra closely match the phosphorescence of the corresponding carbonyl compounds. The reactivity of both carbenes towards O2 is very different. Singlet carbene phenylchlorocarbene reacts thermally only slowly with O2 up to 60 K. The oxidation products phenylchloroformate, benzoyl chloride and O(3P) are mainly formed photochemically on irradiation of the diazirine precursor. Triplet carbene (2-chlorophenyl)carbene reacts readily with O2 at cryogenic temperatures to give mostly 2-chlorobenzaldehyde-O-oxide. The carbonyl-O-oxide is photochemically easily cleaved to give 2-chlorobenzaldehyde and O(3P). The reaction step leading to carbonyl compounds in their excited states is in both carbene oxidations the recombination of the free carbene and O(3P).  相似文献   

12.
The gas‐phase reaction of monomethylhydrazine (CH3NH? NH2; MMH) with ozone was investigated in a flow tube at atmospheric pressure and a temperature of 295 ± 2 K using N2/O2 mixtures (3–30 vol% O2) as the carrier gas. Proton transfer reaction–mass spectrometry (PTR‐MS) and long‐path FT‐IR spectroscopy served as the main analytical techniques. The kinetics of the title reaction was investigated with a relative rate technique yielding kMMH+O3 = (4.3 ± 1.0) × 10?15 cm3 molecule?1 s?1. Methyldiazene (CH3N?NH; MeDia) has been identified as the main product in this reaction system as a result of PTR‐MS analysis. The reactivity of MeDia toward ozone was estimated relative to the reaction of MMH with ozone resulting in kMeDia+O3 = (2.7 ± 1.6) × 10?15 cm3 molecule?1 s?1. OH radicals were followed indirectly by phenol formation from the reaction of OH radicals with benzene. Increasing OH radical yields with increasing MMH conversion have been observed pointing to the importance of secondary processes for OH radical generation. Generally, the detected OH radical yields were definitely smaller than thought so far. The results of this study do not support the mechanism of OH radical formation from the reaction of MMH with ozone as proposed in the literature.  相似文献   

13.
The reaction rate of the coulometric variant of the Karl-Fischer titration reaction (in which electrolytically generated triiodide is used as oxidant instead of iodine) has been measured in methanol. The reaction is first order in water, sulfur dioxide and triiodide, respectively. For pH<5 the reaction rate constant decreases logarithmically with decreasing pH. Addition of pyridine solely influences the pH (by fixing it to a value of about 6) and has no direct influence on the reaction rate. A linear relation exists between the reaction rate constant and the reciprocal value of the iodide concentration, from which we can calculate the individual reaction rates for the oxidation by iodine and triiodide, respectively. While the reaction rate constant for triiodide is relatively small (k3≈350 l2 mol?2s?1), the reaction rate constant for iodine is much larger (k3≈1.5×107 l2 mol?2 s?1.  相似文献   

14.
An unexpected approach to the preparation of quinoxaline and pyrido[2,3‐b]pyrazine derivatives 5 is described. The reaction between 1H‐indole‐2,3‐diones 1 , 1‐phenyl‐2‐(triphenylphosphoranylidene)ethanone ( 2 ), and benzene‐1,2‐ or pyridine‐2,3‐diamines 3 proceeds in MeOH under reflux in good to excellent yields (Scheme 1 and Table). No co‐catalyst or activator is required for this multi‐component reaction (MCR), and the reaction is, from an experimental point of view, simple to perform. The structures of 5, 5′ , and 6 were corroborated spectroscopically (IR, 1H‐ and 13C‐NMR, and EI‐MS) and were confirmed by comparison with reference compounds. A plausible mechanism for this type of reaction is proposed (Scheme 2).  相似文献   

15.
Cyanoaromatic sensitizers, in particular 9,10-dicyanoanthracene (DCA), sensitize the photo-oxygenation of olefins by two distinct mechanisms. In the case of aryl substituted olefins (OL), which react extremely slowly (if at all) with singlet oxygen, the reaction proceeds by way of electron transfer to produce discrete radical ions (DCA-and OL+). In the presence of oxygen, this ionic process results, ultimately, in the cleavage of the olefin to carbonyl compounds along with production of some epoxide and other minor by-products. Aromatic ethers can interfere with this process by reducing the radical cation by electron transfer, resulting in net quenching of the reaction. With simple alkenes the DCA-sensitized reaction takes a different course, producing hydroperoxide products with distributions which are very similar to those obtained with the singlet oxygen ene reaction. Careful study has shown that this reaction does, indeed, proceed by way of singlet oxygen, which is produced by at least two mechanisms : (1) enhanced intersystem crossing, in which 1DCA is quenched by interaction with the olefin, leading to a low yield of 3DCA, which subsequently reacts with oxygen to produce singlet oxygen; and (2) direct reaction of 1DCA with oxygen. At limiting high oxygen concentration, this process produces 2 mol of singlet oxygen for each mol of 1DCA quenched; the mechanism involves energy transfer to produce 3DCA and 1 mol of singlet oxygen ; the 3DCA reacts again with oxygen to produce a second mol of singlet oxygen. The complex kinetic behaviour of simple olefins in the presence of DCA can be satisfactorily rationalized by these mechanisms.  相似文献   

16.
The Belousov‐Zhabotinsky (BZ) reaction of bromate ion with 2‐ketoglutaric acid (KGA) in aqueous sulfuric acid catalyzed by Ce(III), Mn(II), or Fe(phen)32+ ion exhibits sustained barely damped oscillations under aerobic conditions. In general, the reaction oscillates without an induction period. Fe(phen)32+ ion behaves differently from Ce(III) and Mn(II) ions in catalyzing this oscillating system. The gem‐diol form of KGA exhibits different behavior from that of the keto form of KGA in the BZ reaction. The kinetics and mechanism of the reaction of KGA with Ce(IV), Mn(III), or Fe(phen)33+ ion was investigated. The order of relative reactivities of metal ions toward reaction with KGA is Mn(III) > Ce(IV) ≫ Fe(phen)33+. Experimental results are rationalized. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 101–107, 2001  相似文献   

17.
Protonation plays an important role in the redox reactions. We observed this leading role during the reduction of [FeIII(phen)2(CN)2]+ by FcCOMe and FcCHOHMe. The kinetic data showed that the reaction(s) followed a complex kinetics due to the formation of protonated acetylferrocene (FcC+OHMe), and or, protonated α-methylferrocenemethanol (FcCHO+H2Me) in aqueous dioxane (80% v/v). Our results helped us to conclude that the reactions were completed in three phases. An overall zeroth order was observed in the first phase of the reactions. In the second phase, the kinetic data showed an overall second order reaction. The third phase was a complex phase where the rate of redox reactions and the insolubility of the neutral product ([FeII(phen)2(CN)2]) competed with each other. We studied the effect of different factors to identify the reacting entities, which take part in the rate-determining step of each reaction in the second phase. Consequently, we determined the effects of selected factors upon the observed pseudo-first order rate constant(s) (k′ obs) of each reaction. The value of k′ obs increased upon addition of protons in the reaction mixture in case of FcCOMe, and it decreased during the oxidation of FcCHOHMe. Meanwhile, upon enhancing the ionic strength, we observed an increase in k′ obs for FcCOMe, and no change in the value of k′ obs during the reaction of FcCHOHMe. However, a decrease in k′ obs was noticed upon increasing the dielectric constant of the reaction mixture when the reductant was FcCOMe, and no effect was observed in case of FcCHOHMe. Together, these results suggested oxidation of FcC+OHMe and FcCHOHMe in the slow-step, and FcCOMe and FcCHO+H2Me during the fast-step. We refined our results by estimating the thermodynamic parameters of activation. The low values of activation energy and enthalpy of activation confirmed that the reduction of [FeIII(phen)2(CN)2]+ hardly depends upon temperature when the reducing agent is FcCOMe. The outcomes justified that the rate of reaction depends upon the unsaturated FcC+OHMe. This intermediate species contain a ‘carbonium ion’, which is very reactive and energetic. We obtained comparatively high values of the activation energy and enthalpy of activation for the reaction between [FeIII(phen)2(CN)2]+ and FcCHOHMe. The results show that FcCHOHMe is a saturated and stable compound that leads the slow-step and controls the rate of reaction.  相似文献   

18.
Formic acid (HCOOH, FA) has long been considered as a promising hydrogen-storage material due to its efficient hydrogen release under mild conditions. In this work, FA decomposes to generate CO2 and H2 selectively in the presence of aqueous Pd2+ complex solutions at 333 K. Pd(NO3)2 was the most effective in generating H2 among various Pd2+ complexes explored. Pd2+ complexes were in situ reduced to Pd0 species by the mixture of FA and sodium formate (SF) during the course of the reaction. Since C−H activation reaction of Pd2+-bound formate is occurred for both Pd2+ reduction and H2/CO2 gas generation, FA decomposition pathways using several Pd2+ species were explored using density functional theory (DFT) calculations. Rotation of formate bound to Pd2+, β-hydride elimination, and subsequent CO2 and H2 elimination by formic acid were examined, providing different energies for rate determining step depending on the ligand electronics and geometries coordinated to the Pd2+ complexes. Finally, Pd2+ reduction toward Pd0 pathways were explored computationally either by generated H2 or reductive elimination of CO2 and H2 gas.  相似文献   

19.
The O(3P) + C2H2 reaction plays an important role in hydrocarbon combustion. It has two primary competing channels: H + HCCO (ketenyl) and CO + CH2 (triplet methylene). To further understand the microscopic dynamic mechanism of this reaction, we report here a detailed quasi-classical trajectory study of the O(3P) + C2H2 reaction on the recently developed full-dimensional potential energy surface (PES). The entrance barrier TS1 is the rate-limiting barrier in the reaction. The translation of reactants can greatly promote reactivity, due to strong coupling with the reaction coordinate at TS1. The O(3P) + C2H2 reaction progress through a complex-forming mechanism, in which the intermediate HCCHO lives at least through the duration of a rotational period. The energy redistribution takes place during the creation of the long-lived high vibrationally (and rotationally) excited HCCHO in the reaction. The product energy partitioning of the two channels and CO vibrational distributions agree with experimental data, and the vibrational state distributions of all modes of products present a Boltzmann-like distribution.  相似文献   

20.
Density functional theory (DFT) calculations were used to study the mechanism of CO2 hydrolysis by Zn‐(1,5,9‐triazacyclododecane) and Zn‐cyclam and evaluate the associated thermodynamic and kinetic parameters. Microkinetic models were then built based on the kinetics and thermodynamics derived from first principles. Both catalysts showed very similar behavior to Zn‐cyclen, which we have reported previously, but with multiple distinctions. The intrinsic reaction rate constants for Zn‐(1,5,9‐triazacyclododecane) and Zn‐cyclam were calculated to be 2693 and 4623 M?1 s?1, respectively, which is in reasonable agreement with experimental values reported or estimated. The CO2 adsorption step was found to be a rate‐limiting step for all three catalysts. Zn‐cyclam has the lowest barrier for this step due to the highest pKa or nucleophilicity of the Zn‐OH? form, and, therefore, the highest intrinsic activity. However, the observed reaction rate constant also depends on the availability of the catalyst. The decrease in the observed reaction rate constant over 0–12 ms was ascribed to the decrease in the concentration of the catalytic form, Zn‐OH?, which was primarily converted to Zn‐HCO3?. The reaction rate constant of Zn‐cyclam dropped much faster than those of Zn‐cyclen and Zn‐(1,5,9‐triazacyclododecane) due to lower energy of the Zn‐HCO3? form. The conversion of CO2 at 1000 ms as a function of pH was calculated to compare the relative activity of these catalysts, and Zn‐cyclen was found to be the best catalyst.  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