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 mechanism of the dirhodium tetracarboxylate catalyzed cyclopropanation of alkenes with both unsubstituted diazoacetates and vinyl- and phenyldiazoacetates was studied by a combination of (13)C kinetic isotope effects and density functional theory calculations. The cyclopropanation of styrene with methyl phenyldiazoacetate catalyzed by Rh(2)(octanoate)(4) exhibits a substantial (13)C isotope effect (1.024) at the terminal olefinic carbon and a smaller isotope effect (1.003-1.004) at the internal olefinic carbon. This is consistent with a highly asynchronous cyclopropanation process. Very similar isotope effects were observed in a bisrhodium tetrakis[(S)-N-(dodecylbenzenesulfonyl)prolinate] (Rh(2)(S-DOSP)(4) catalyzed reaction, suggesting that the chiral catalyst engages in a very similar cyclopropanation transition-state geometry. Cyclopropanation with ethyl diazoacetate was concluded to involve an earlier transition state, based on a smaller terminal olefinic isotope effect (1.012-1.015). Density functional theory calculations (B3LYP) predict a reaction pathway involving complexation of the diazoesters to rhodium, loss of N(2) to afford a rhodium carbenoid, and an asynchronous but concerted cyclopropanation transition state. The isotope effects predicted for reaction of a phenyl-substituted rhodium carbenoid with styrene match within the error of the experimental values, supporting the accuracy of the theoretical calculations and the rhodium carbenoid mechanism. The accuracy of the calculations is additionally supported by excellent predictions of reaction barriers, stereoselectivity, and reactivity trends. The nature of alkene selectivity and diastereoselectivity effects in these reactions is discussed, and a new model for enantioselectivity in Rh(2)(S-DOSP)(4)-catalyzed cyclopropanations is presented.  相似文献   

2.
A novel chiral Rh(II) catalyst (1) is introduced for the [2 + 1]-cycloaddition of ethyl diazoacetate to terminal acetylenes and olefins with high enantioselectivity. The catalyst 1 consists of one acetate bridging group and three mono-N-triflyldiphenylimidazoline-2-one bidentate ligands (DPTI) spanning the Rh(II)-Rh(II) metallic center in a structure that was determined by single-crystal X-ray diffraction analysis. A rational mechanism is advanced that provides a straightforward explanation for the enantioselectivity and absolute stereochemical course of the [2 + 1]-cycloaddition reactions. A key element in this explanation is the cleavage of one of the Rh-O bonds of the bridging acetate group in the intermediate Rh-carbene complex to form a new pentacoordinate Rh carbene complex (formally 1.5 valent Rh) that can undergo [2 + 2]-cycloaddition with the C-C pi-bond of the acetylenic or olefinic substrate. Reductive elimination of the resulting adduct affords the cyclopropene or cyclopropane product. The C2-symmetry of the two DPTI ligands orthogonal to the bridging acetate also contributes to the high observed enantioselectivity and mechanistic clarity. The catalyst 1, which functions effectively at 0.5 mol %, can be recovered efficiently for reuse. Its ready availability, robustness, and effectiveness suggest it as a useful addition to the list of practical chiral Rh(II) catalysts for synthesis.  相似文献   

3.
Carbon-13 kinetic isotope effects (KIEs) have been determined for free-radical and copper-mediated living radical polymerizations of methyl methacrylate at 60 degrees C. While free-radical polymerization shows only one primary 13C KIE, on the least-substituted double bond carbon (k12/k13 = 1.045), two significant KIEs are observed, one on each double bond carbon, for copper-mediated polymerization (k12/k13(H2C=) = 1.050, k12/k13(=C <) = 1.010), showing that copper-mediated living radical polymerization does not propagate via a simple free radical process.  相似文献   

4.
The diastereoselectivity of the Rh-catalyzed C-H amination was examined with 18 chiral open-chain substrates, which bear a benzylic methylene group in the α-position to a stereogenic center (-CHMeX), and with four chiral cyclic tetralins, in which the stereogenic center was positioned at carbon atom C2. The C-H amination was performed using trichloroethoxysulfonyl-substituted amine (H(2)NTces) as the nitrogen source, a diacyloxyiodobenzene as the oxidant, and bis[rhodium(α,α,α',α'-tetramethyl-1,3-benzenedipropionate)] [Rh(2)(esp)(2)] as the catalyst. For acyclic substrates a high syn diastereoselectivity (dr > 95/5) was found if the substituent X was Br, PO(OEt)(2), SO(2)Ph, or OOCCF(3) (eight examples). Moderate to good syn selectivities (dr = 80/20 to 91/9) were found for X = NO(2), OAc, COOMe, and CN (eight examples). Only two substrates gave a low diastereoselectivity. Kinetic isotope effect (KIE) experiments revealed that there is no secondary KIE when replacing -CHMeCOOMe by -CDMeCOOMe, but there is a significant primary KIE at the benzylic methylene position (4.8 ± 0.7). Deuteration experiments provided evidence that the reaction proceeds stereospecifically with retention of configuration. A preferred conformation is proposed, which explains the outcome of the reaction. In this conformation the X substituent is antiperiplanar to the C-H bond, which is diastereoselectively attacked, and steric strain between the remaining substituents at the stereogenic and the prostereogenic center is minimized. DFT calculations support this model. They suggest, however, that the reaction is not concerted but occurs via hydrogen atom abstraction and subsequent radical rebound. Further support for an antiperiplanar attack relative to a given substituent X = Br, COOMe, or CN was obtained with the respective 2-substituted tetralins. Attack at C1 provides almost exclusively the trans-amination product. If the size of the X substituent increases [Br < CN < COOMe < PO(OEt)(2)], attack at the carbon atom C4 prevails, delivering the respective trans-amination products at this position.  相似文献   

5.
Kinetic isotope effects (KIEs) were measured for methyl glucoside (4) hydrolysis on unlabeled material by NMR. Twenty-eight (13)C KIEs were measured on the acid-catalyzed hydrolysis of alpha-4 and beta-4, as well as enzymatic hydrolyses with yeast alpha-glucosidase and almond beta-glucosidase. The 1-(13)C KIEs on the acid-catalyzed reactions of alpha-4 and beta-4, 1.007(2) and 1.010(6), respectively, were in excellent agreement with the previously reported values (1.007(1), 1.011(2): Bennet and Sinnott, J. Am. Chem. Soc. 1986, 108, 7287). Transition state analysis of the acid-catalyzed reactions using the (13)C KIEs, along with the previously reported (2)H KIEs, confirmed that both reactions proceed with a stepwise D(N)A(N) mechanism and showed that the glucosyl oxocarbenium ion intermediate exists in an E(3) sofa or (4)H(3) half-chair conformation. (13)C KIEs showed that the alpha-glucosidase reaction also proceeded through a D(N)*A(N) mechanism, with a 1-(13)C KIE of 1.010(4). The secondary (13)C KIEs showed evidence of distortions in the glucosyl ring at the transition state. For the beta-glucosidase-catalyzed reaction, the 1-(13)C KIE of 1.032(1) demonstrated a concerted A(N)D(N) mechanism. The pattern of secondary (13)C KIEs was similar to the acid-catalyzed reaction, showing no signs of distortion. KIE measurement at natural abundance makes it possible to determine KIEs much more quickly than previously, both by increasing the speed of KIE measurement and by obviating the need for synthesis of isotopically labeled compounds.  相似文献   

6.
Insight into the N7/O6 equatorial binding interactions of the antitumor active complex Rh(2)(OAc)(4)(H(2)O)(2) (OAc(-) = CH(3)CO(2)(-)) with the nucleotide 5'-GMP and the DNA fragment d(pGpG) has been obtained by one- (1D) and two-dimensional (2D) NMR spectroscopy. The lack of N7 protonation at low pH values and the significant increase in the acidity of N1-H (pK(a) approximately 5.6 as compared to 8.5 for N7 only bound platinum adducts), indicated by the pH dependence study of the H8 (1)H NMR resonance for the HT (head-to-tail) isomer of Rh(2)(OAc)(2)(5'-GMP)(2), are consistent with bidentate N7/O6 binding of the guanine. The H8 (1)H NMR resonance of the HH (head-to-head) Rh(2)(OAc)(2)(5'-GMP)(2) isomer, as well as the 5'-G and 3'-G H8 resonances of the Rh(2)(OAc)(2) [d(pGpG)] adduct exhibit pH-independent titration curves, attributable to the added effect of the 5'-phosphate group deprotonation at a pH value similar to that of the N1 site. The enhancement in the acidity of N1-H, with respect to N7 only bound metal adducts, afforded by the O6 binding of the bases to the rhodium centers, has been corroborated by monitoring the pH dependence of the purine C6 and C2 (13)C NMR resonances for Rh(2)(OAc)(2)(5'-GMP)(2) and Rh(2)(OAc)(2) [d(pGpG)]. The latter studies resulted in pK(a) values in good agreement with those derived from the pH-dependent (1)H NMR titrations of the H8 resonances. Comparison of the (13)C NMR resonances of C6 and C2 for the dirhodium adducts Rh(2)(OAc)(2)(5'-GMP)(2) and Rh(2)(OAc)(2) [d(pGpG)] with the corresponding resonances of the unbound ligands at pH 8.0, showed substantial downfield shifts of Deltadelta approximately 11.0 and 6.0 ppm, respectively. The HH arrangement of the bases in the Rh(2)(OAc)(2) [d(pGpG)] adduct is evidenced by intense H8/H8 ROE cross-peaks in the 2D ROESY NMR spectrum. The presence of the terminal 5'-phosphate group in d(pGpG) results in stabilization of one left-handed Rh(2)(OAc)(2) [d(pGpG)] HH1 L conformer, due to the steric effect of the 5'-group, favoring left canting in cisplatin-DNA adducts. Complete characterization of the Rh(2)(OAc)(2[d(pGpG)] adduct revealed notable structural features that resemble those of cis-[Pt(NH(3))(2) [d(pGpG)]]; the latter involve repuckering of the 5'-G sugar ring to C3'-endo (N-type) conformation, retention of C2'-endo (S-type) 3'-G sugar ring conformation, and anti orientation with respect to the glycosyl bonds. The superposition of the low energy Rh(2)(OAc)(2) [d(pGpG)] conformers, generated by simulated annealing calculations, with the crystal structure of cis-[Pt(NH(3))(2) [d(pGpG)]], reveals remarkable similarities between the adducts; not only are the bases almost completely destacked upon coordination to the metal in both cases, but they are favorably poised to accommodate the bidentate N7/O6 binding to the dirhodium unit. Unexpectedly, the two metal-metal bonded rhodium centers are capable of engaging in cis binding to GG intrastrand sites by establishing N7/O6 bridges that span the Rh-Rh bond.  相似文献   

7.
Multiple isotope effects were measured at the reactive center of formamide during acid-catalyzed hydrolysis in water at 25 degrees C. The mechanism involves a rapid pre-equilibrium protonation of the carbonyl oxygen, followed by the formation of at least one tetrahedral intermediate, which does not appreciably exchange its carbonyl oxygen with the solvent (kh/kex = 55). The pKa for formamide was determined by 15N NMR and found to be about -2.0. The formyl-hydrogen kinetic isotope effect (KIE) is indicative of a transition state that is highly tetrahedral (Dkobs = 0.79); the carbonyl-carbon KIE (13kobs = 1.031) is in agreement with this conclusion. The small leaving-nitrogen KIE (15kobs = 1.0050) is consistent with some step prior to breaking the C-N bond as rate-determining. The carbonyl-oxygen KIE (18kobs = 0.996) points to attack of water as the rate-determining step. On the basis of these results, a mechanism is proposed in which attachment of the nucleophile to a protonated formamide molecule is rate determining.  相似文献   

8.
A mechanistic study of the Cp*Rh(III)-catalyzed annulative coupling of benzimidates with 4-acyl-1-sulfonyltriazoles by C H activation was performed using density functional M06 method. It was demonstrated that the active catalyst during the coupling process should be the cation [Cp*Rh(OAc)]+ rather than the neutral Cp*Rh(OAc)2 and Zn(OAc)2 as proposed previously by the experimenters. A novel energetically feasible reaction pathway has been revealed theoretically in details. The acetic acid-mediated cyclization process was confirmed to be the rate-limiting step with an overall barrier of 24.3 kcal/mol, excluding any importance of the C H cleavage mechanism as supported by the kinetic isotope effect experiments. The major factors responsible for the preferred regioselectivity of N-pyrimidinylinoles with 4-acetyl-1-sulfonyltriazoles were discussed.  相似文献   

9.
The B3LYP density functional studies on the dirhodium tetracarboxylate-catalyzed C-H bond activation/C-C bond formation reaction of a diazo compound with an alkane revealed the energetics and the geometry of important intermediates and transition states in the catalytic cycle. The reaction is initiated by complexation between the rhodium catalyst and the diazo compound. Driven by the back-donation from the Rh 4d(xz) orbital to the C[bond]N sigma*-orbital, nitrogen extrusion takes place to afford a rhodium[bond]carbene complex. The carbene carbon of the complex is strongly electrophilic because of its vacant 2p orbital. The C[bond]H activation/C[bond]C formation proceeds in a single step through a three-centered hydride transfer-like transition state with a small activation energy. Only one of the two rhodium atoms works as a carbene binding site throughout the reaction, and the other rhodium atom assists the C[bond]H insertion reaction. The second Rh atom acts as a mobile ligand for the first one to enhance the electrophilicity of the carbene moiety and to facilitate the cleavage of the rhodium[bond]carbon bond. The calculations reproduce experimental data including the activation enthalpy of the nitrogen extrusion, the kinetic isotope effect of the C[bond]H insertion, and the reactivity order of the C[bond]H bond.  相似文献   

10.
Mononuclear rhodium complexes with reactive olefin ligands, supported on MgO powder, were synthesized by chemisorption of Rh(C(2)H(4))(2)(C(5)H(7)O(2)) and characterized by infrared (IR), (13)C MAS NMR, and extended X-ray absorption fine structure (EXAFS) spectroscopies. IR spectra show that the precursor adsorbed on MgO with dissociation of acetylacetonate ligand from rhodium, with the ethylene ligands remaining bound to the rhodium, as confirmed by the NMR spectra. EXAFS spectra give no evidence of Rh-Rh contributions, indicating that site-isolated mononuclear rhodium species formed on the support. The EXAFS data also show that the mononuclear complex was bonded to the support by two Rh-O bonds, at a distance of 2.18 A, which is typical of group 8 metals bonded to oxide supports. This is the first simple and nearly uniform supported mononuclear rhodium-olefin complex, and it appears to be a close analogue of molecular catalysts for olefin hydrogenation in solution. Correspondingly, the ethylene ligands bonded to rhodium in the supported complex were observed to react with H(2) to form ethane, and the supported complex was catalytically active for the ethylene hydrogenation at 298 K. The ethylene ligands also underwent facile exchange with C(2)D(4), and exposure of the sample to carbon monoxide led to the formation of rhodium gem dicarbonyls.  相似文献   

11.
Complexation of the oxygen atom in 2-butyl phenyl ethers to a rhodium atom of the dirhodium tetracarboxylate Rh(II) 2[(R)-(+)-MTPA]4(Rh*, MTPA-H = methoxytrifluoromethylphenylacetic acid identical with Mosher's acid) deshields an sp3-hybridized 13C nucleus directly bonded to the ether oxygen; apparently, the inductive effect of the oxygen is enhanced when it is complexed to the rhodium atom. On the other hand, deshielding complexation shifts of aromatic ipso-carbons (alpha-positioned) are minute but ortho- and para-carbon signals are influenced by the resonance effect of oxygen. This effect can be modulated by further substituents at the benzene ring. In turn, this modulation of the resonance correlates linearly ith the magnitude of the inductive effect exerted on the aliphatic alpha-carbon atoms. Diastereomeric dispersion effects at 13C signals can be observed for most compounds, indicating that enantiodifferentiation is possible in this class of ethers.  相似文献   

12.
Using 13C NMR techniques, the 12C/13C kinetic isotope effects (KIEs) for the polymerization of 1-hexene catalyzed by rac-(C2H4(1-indenyl)2)ZrMe2 in the presence of four different cocatalysts (tris(pentfluorophenyl)borane, tris(pentafluorophenyl)alane, anilinium tetrakis(pentafluorophenyl)borate, and methylalumoxane) have been determined. All cocatalysts yield similar KIEs within experimental uncertainty, with values of 1.009(1) and 1.017(2) at C1 and C2, respectively. Ab initio DFT computational modeling of the polymerization KIE indicates that alkene binding to the catalyst must be reversible, with the majority of the KIE developing in the subsequent migratory insertion reaction.  相似文献   

13.
Ribulose-1,5-bisphosphate carboxylase/oxygenase (RubisCO) fixes atmospheric carbon dioxide into bioavailable sugar molecules. It is also well known that a kinetic isotope effect (KIE; CO2 carbon atoms) accompanies the carboxylation process. To describe the reaction and the KIE α, two different types of molecular dynamics (MD) simulations (ab initio MD and classical MD) have been performed with an Own N-layered Integrated molecular Orbitals and molecular Mechanics (ONIOM)-hybrid model. A channel structure for CO2 transport has been observed during the MD simulation in RubisCO, and assuming the reaction path from the inlet to the product through the coordinate complex with Mg2+, simulations have been performed on several molecular configuration models fixing several distances between CO2 and ribulose-1,5-bisphosphate along the channel. Free energy analysis and diffusion coefficient analysis have been evaluated for different phases of the process. It is confirmed that the isotopic fractionation effect for CO2 containing either 13C or 12C would appear through the transiting path in the channel structure identified in RubisCO. The estimated isotope fractionation constant was quite close to the experimental value.  相似文献   

14.
Observed pseudo-first-order rate constants (k(obs)) of the hydride-transfer reactions from isopropyl alcohol (i-PrOH) to two NAD(+) analogues, 9-phenylxanthylium ion (PhXn(+)) and 10-methylacridinium ion (MA(+)), were determined at temperatures ranging from 49 to 82 degrees C in i-PrOH containing various amounts of AN or water. Formations of the alcohol-cation ether adducts (ROPr-i) were observed as side equilibria. The equilibrium constants for the conversion of PhXn(+) to PhXnOPr-i in i-PrOH/AN (v/v = 1) were determined, and the equilibrium isotope effect (EIE = K(i-PrOH)/K(i-PrOD)) at 62 degrees C was calculated to be 2.67. The k(H) of the hydride-transfer step for both reactions were calculated on the basis of the k(obs) and K. The corresponding deuterium kinetic isotope effects (e.g., KIE(OD)(H) = k(H)(i-PrOH)/k(H)(i-PrOD) and KIE(beta-D6)(H) = k(obs)(i-PrOH)/k(obs)((CD3)2CHOH)), as well as the activation parameters, were derived. For the reaction of PhXn(+) (62 degrees C) and MA(+) (67 degrees C), primary KIE(alpha-D)(H) (4.4 and 2.1, respectively) as well as secondary KIE(OD)(H) (1.07 and 1.18) and KIE(beta-D6)(H) (1.1 and 1.5) were observed. The observed EIE and KIE(OD)(H) were explained in terms of the fractionation factors for deuterium between OH and OH(+)(OH(delta+)) sites. The observed inverse kinetic solvent isotope effect for the reaction of PhXn(+) (k(obs)(i-PrOH)/k(obs)(i-PrOD) = 0.39) is consistent with the intermolecular hydride-transfer mechanism. The dramatic reduction of the reaction rate for MA(+), when the water or i-PrOH cosolvent was replaced by AN, suggests that the hydride-transfer T.S. is stabilized by H-bonding between O of the solvent OH and the substrate alcohol OH(delta+). This result suggests an H-bonding stabilization effect on the T.S. of the alcohol dehydrogenase reactions.  相似文献   

15.
The N7/O6 equatorial binding interactions of the antitumor active complex Rh(2)(OAc)(4)(H(2)O)(2) (OAc(-) = CH(3)CO(2)(-)) with the DNA fragment d(GpG) have been unambiguously determined by NMR spectroscopy. Previous X-ray crystallographic determinations of the head-to-head (HH) and head-to-tail (HT) adducts of dirhodium tetraacetate with 9-ethylguanine (9-EtGH) revealed unprecedented bridging N7/O6 guanine nucleobases that span the Rh-Rh bond. The absence of N7 protonation at low pH and the notable increase in the acidity of N1-H (pK(a) approximately 5.7 as compared to 8.5 for N7 only bound platinum adducts), suggested by the pH dependence titrations of the purine H8 (1)H NMR resonances for Rh(2)(OAc)(2)(9-EtG)(2) and Rh(2)(OAc)(2-)[d(GpG)],are consistent with bidentate N7/O6 binding of the guanine nucleobases. The pK(a) values estimated for N1-H (de)protonation, from the pH dependence studies of the C6 and C2 (13)C NMR resonances for the Rh(2)(OAc)(2)(9-EtG)(2) isomers, concur with those derived from the H8 (1)H NMR resonance titrations. Comparison of the (13)C NMR resonances of C6 and C2 for the dirhodium adducts Rh(2)(OAc)(2)(9-EtG)(2) and Rh(2)(OAc)(2)[d(GpG)] with the corresponding resonances of the unbound ligands [at pH 7.0 for 9-EtGH and pH 8.0 for d(GpG)], shows substantial downfield shifts of Deltadelta approximately 11.0 and 6.0 ppm for C6 and C2, respectively; the latter shifts reflect the effect of O6 binding to the dirhodium centers and the ensuing enhancement in the acidity of N1-H. Intense H8/H8 ROE cross-peaks in the 2D ROESY NMR spectrum of Rh(2)(OAc)(2)[d(GpG)] indicate head-to-head arrangement of the guanine bases. The Rh(2)(OAc)(2)[d(GpG)] adduct exhibits two major right-handed conformers, HH1 R and HH2 R, with HH1 R being three times more abundant than the unusual HH2 R. Complete characterization of both adducts revealed repuckering of the 5'-G sugar rings to C3'-endo (N-type), retention of C2'-endo (S-type) conformation for the 3'-G sugar rings, and anti orientation with respect to the glycosyl bonds. The structural features obtained for Rh(2)(OAc)(2))[d(GpG)] by means of NMR spectroscopy are very similar to those for cis-[Pt(NH(3))(2))[d(GpG)]] and corroborate molecular modeling studies.  相似文献   

16.
Nishino F  Miki K  Kato Y  Ohe K  Uemura S 《Organic letters》2003,5(15):2615-2617
[reaction: see text] The reaction of alkenes with conjugated ene-yne-imino ether or ene-yne-aldimine in the presence of a catalytic amount of [Rh(OAc)(2)](2) gives (2-pyrrolyl)cyclopropanes in good yields. The key intermediate of this cyclopropanation is a (2-pyrrolyl)carbenoid generated by the nucleophilic attack of imine nitrogen atom at an internal alkyne carbon activated by rhodium complex. The intramolecular reaction also proceeds to afford a polycyclic pyrrole.  相似文献   

17.
The binuclear rhodium(II) complex Na2[Rh2Cl2(OAc)4]· 4H2O has an infinite sheet structure. Binuclear anionic complexes [Rh2Cl2(OAc)4]2– are bound with cationic entities. The Na+ cation has pseudooctahedral coordination and is surrounded by two chloro ligands and two oxygen atoms of bridging acetato ligands of two [Rh2Cl2(OAc)4]2– anions and two water molecules. Both Cl and H2O are bridging ligands involved in formation of the Na+ chains. The remaining water molecules are located between sheets.  相似文献   

18.
A rhodium catalyst supported on a Sibunit graphitized carbon carrier was studied by in situ XAFS spectroscopy. A comparative study of the reduction of rhodium was performed for the following two samples: Rh/C(120) dried at 120°C and Rh/C(350) calcined at 350°C. EXAFS data showed an absence of carbon atoms within the nearest environment of rhodium atoms in the Rh/C(120) uncalcined sample, which implies the absence of direct interaction between rhodium and the carbon support. In the course of the reduction of this sample (200°C), coarse particles with small metal cores were initially formed. These metal particles rapidly agglomerated upon the complete reduction of rhodium (350°C). These reduction of the Rh/C(350) calcined sample at 100–500°C resulted in the formation of small metal particles early in the reduction (100°C). The high dispersity of these particles was retained as the temperature of treatment in hydrogen was increased to 500°C due to metal-support interaction. The conversion of benzene into cyclohexane on the Rh/C(350) catalyst containing small rhodium particles was much higher at the same temperature of hydrogenation.Translated from Kinetika i Kataliz, Vol. 46, No. 1, 2005, pp. 122–130. Original Russian Text Copyright © 2005 by Stakheev, Tkachenko, Klementev, Grünert, Bragina, Mashkovskii, Kustov.  相似文献   

19.
The kinetic isotope effect (KIE) for carbon and oxygen in the reaction CO + OH has been measured over a range of pressures of air and at 0.2 and 1.0 atm of oxygen, argon, and helium. The reaction was carried out with 21–86% conversion under static conditions, utilizing the photolysis of H2O2 as a source of OH radicals. The value of the KIE for carbon varies with pressure and the kind of ambient gas; for air the ratio of the reaction rates 12k/13k has the value 1.007 at 1.00 atm and decreases to 0.997 at 0.2 atm; for oxygen and argon over the same pressure range the values are 1.002–0.994 and 1.000–0.991, respectively. The value of the KIE for the CO oxygen atom is 16k/18k = 0.990 over the pressure range 0.2–1.0 atm and is independent of the kind of ambient gas. No exchange of the oxygen atoms in the activated complex, followed by decomposition to the starting molecules, was observed. From the mechanistic standpoint the normal KIE observed for carbon at the high pressure is attributed to the initial formation of the activated HOCO radical, whereas the inverse KIE observed at low pressures is a result of the KIE for the reverse reaction HOCO? → CO + OH being greater than that for the forward reaction HOCO? → CO2 + H. The derived isotopic equilibrium constant for HOCO ?CO favors the enrichment of 13C in the more strongly bound HOCO.  相似文献   

20.
Rhodium transition-metal-organic cooperative catalysis, which has been intensively studied by many chemists, represents a great success in C–H bond activation because of high efficiencies and selectivities. Typically, in the reaction mechanism of aldehyde and alkene catalyzed by Rh(I) complex and 2-amino-3-picoline, two kinds of metala-cyclic transition-metal complexes of (iminoacyl)rhodium(III)hydride and (iminoacyl)rhodium(III) alkyl are generally formed. The two complexes play an important role in the overall reaction, in which the Rh–C bond formations are involved. So it is meaningful to understand the strength of Rh–C bond, which can be measured by the homolytic bond dissociation enthalpies (BDEs). To this end, we first calculated 16 relative Rh–C BDEs of Tp′Rh(CNneopentyl)RH (Tp′?=?hydridotris-(3,5-dimethylpyrazolyl)borate) by 19 density functional theory (DFT) methods. Furthermore, the 5 absolute Rh–C BDEs of Rh transition-metal complexes were also calculated. The results show that the B97D3 is the most accurate method to predict the relative and absolute Rh–C BDEs and the corresponding RMSE values are the smallest of 2.8 and 3.3?kcal/mol respectively. Therefore, the Rh–C BDEs of (iminoacyl)rhodium(III)hydride and (iminoacyl)rhodium(III)alkyl as well as the substituent effects were investigated by using the B97D3 method. The results indicated that the different substituents exhibit different effects on different types of Rh–C BDEs. In addition, the analysis including the natural bond orbital (NBO) as well as the energies of frontier orbitals were performed in order to further understand the essence of the Rh–C BDE change patterns.  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