Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 28
Filter
Add more filters










Publication year range
1.
J Org Chem ; 78(14): 7194-8, 2013 Jul 19.
Article in English | MEDLINE | ID: mdl-23829207

ABSTRACT

The mechanism of base-catalyzed rearrangement of ring-substituted benzoins in aqueous methanol was examined by kinetic and product analyses. Substituent effects on the rate and equilibrium constants revealed that the kinetic process has a different electron demand compared to the equilibrium process. Reactions in deuterated solvents showed that the rate of H/D exchange of the α-hydrogen is similar to the overall rate toward the equilibrium state. A proton-inventory experiment using partially deuterated solvents showed a linear dependence of the rate on the deuterium fraction of the solvent, indicating that only one deuterium isotope effect contributes to the overall rate process. All these results point to a mechanism in which the rearrangement is initiated by the rate-determining α-hydrogen abstraction rather than a mechanism with initial hydroxyl hydrogen abstraction as in the general α-ketol rearrangement.


Subject(s)
Benzoin/chemistry , Hydroxides/chemistry , Potassium Compounds/chemistry , Catalysis , Kinetics , Methanol/chemistry , Molecular Structure
2.
J Org Chem ; 77(23): 10738-44, 2012 Dec 07.
Article in English | MEDLINE | ID: mdl-23136916

ABSTRACT

The reactions of nitronates of ring-substituted phenylnitromethanes and enolates of ring-substituted 1-phenyl-2-propanones with MeOBs gave exclusively the O-methylated and C-methylated products, respectively. DFT calculations suggested that two factors, namely, intrinsic barriers and metal-cation coordination, control the C/O selectivity. The kinetic preference for O-methylation in the reactions of nitronates arises from the intrinsic barriers, which are ca. 10 kcal/mol lower for O-methylation than for C-methylation. The situation is the same for the gas-phase reaction of an enolate, in which the O-methylation is more favorable than the C-methylation. The experimentally observed C-selectivity of enolate reactions in solution is due to the metal-cation coordination, which hinders O-methylation for enolates. The effects of the enolate reactivity and the solvent on the C/O selectivity are also rationalized to arise from the two factors.

3.
Chemistry ; 18(31): 9714-21, 2012 Jul 27.
Article in English | MEDLINE | ID: mdl-22815219

ABSTRACT

Full-quantum mechanical fragment molecular orbital-based molecular dynamics (FMO-MD) simulations were applied to the hydration reaction of formaldehyde in water solution under neutral conditions. Two mechanisms, a concerted and a stepwise one, were considered with respect to the nucleophilic addition and the proton transfer. Preliminary molecular orbital calculations by means of polarized continuum model reaction field predicted that the hydration prefers a concerted mechanism. Because the calculated activation barriers were too high for free FMO-MD simulations to give reactive trajectories spontaneously, a More O'Ferrall-Jencks-type diagram was constructed from the statistical analysis of the FMO-MD simulations with constraint dynamics. The diagram showed that the hydration proceeds through a zwitterionic-like (ZW-like) structure. The free energy changes along the reaction coordinate calculated by means of the blue moon ensemble for the hydration and the amination of formaldehyde indicated that the hydration proceeds through a concerted process through the ZW-like structure, whereas the amination goes through a stepwise mechanism with a ZW intermediate. In inspection of the FMO-MD trajectories, water-mediated cyclic proton transfers were observed in both reactions on the way from the ZW-like structure to the product. These proton transfers also have an asynchronous character, in which deprotonation from the nucleophilic oxygen atom (or nitrogen atom for amination) precedes the protonation of the carbonyl oxygen atom. The results showed the strong advantage of the FMO-MD simulations to obtain detailed information at a molecular level for solution reactions.

4.
J Org Chem ; 77(8): 4073-8, 2012 Apr 20.
Article in English | MEDLINE | ID: mdl-22486181

ABSTRACT

The Schmidt rearrangement of substituted 3-phenyl-2-butanone with trimethylsilyl azide in 90% (v/v) aqueous TFA gave two types of product, fragmentation and rearrangement, the ratio of which depends on the substituent: more fragmentation for a more electron-donating substituent. Rate measurements by azotometry indicated the presence of an induction period, and the pseudo-first-order rate constants showed saturation kinetics with respect to the azide concentration. It was indicated that the reaction proceeds through pre-equilibrium in the formation of iminodiazonium (ID) ion and that the N(2) liberation from the ID ion is rate-determining. Under high azide concentration conditions, where the effective reactant is the ID ion, the reaction gave a linear Hammett plot with a ρ value of -0.50. The observed substituent effects on the rate and the product selectivity imply that path bifurcation on the way from the rate-determining TS to the product states occurs, as suggested by previous molecular dynamics simulations, in a similar manner to the analogous Beckmann rearrangement/fragmentation reactions.

5.
J Org Chem ; 76(20): 8294-9, 2011 Oct 21.
Article in English | MEDLINE | ID: mdl-21899252

ABSTRACT

The reaction of an α-haloketone with a nucleophile has three reaction channels: carbonyl addition, direct substitution, and proton abstraction. DFT calculations for the reaction of PhCOCH(2)Br with OH(-) showed that there exists an addition/substitution TS on the potential energy surface, in which OH(-) interacts with both the α- and carbonyl carbons. The intrinsic reaction coordinate calculations revealed that the TS serves as the TS for direct substitution for XC(6)H(4)COCH(2)Br with an electron-donating X or a X less electron-withdrawing than m-Cl, whereas the TS serves as the TS for carbonyl addition for derivatives with a X more electron-withdrawing than m-CF(3). Trajectory calculations starting at respective TS indicated that the single TS can serve for the two mechanisms, substitution and addition, through path bifurcation after the TS for borderline substrates. The reaction is the first example of dynamic path bifurcation for fundamental reaction types of carbonyl addition and substitution.

6.
J Org Chem ; 76(11): 4652-60, 2011 Jun 03.
Article in English | MEDLINE | ID: mdl-21563751

ABSTRACT

The reactions of oximes to amides, known as the Beckmann rearrangement, may undergo fragmentation to form carbocations + nitriles when the migrating groups have reasonable stability as cations. The reactions of oxime sulfonates of 1-substituted-phenyl-2-propanone derivatives (7-X) and related substrates (8-X, 9a-X) in aqueous CH(3)CN gave both rearrangement products (amides) and fragmentation products (alcohols), the ratio of which depends on the system; the reactions of 7-X gave amides predominantly, whereas 9a-X yielded alcohols as the major product. The logk-logk plots between the systems gave excellent linear correlations with slopes of near unity. The results support the occurrence of path bifurcation after the rate-determining TS of the Beckmann rearrangement/fragmentation reaction, which has previously been proposed on the basis of molecular dynamics simulations. It was concluded that path-bifurcation phenomenon could be more common than thought and that a reactivity-selectivity argument based on the traditional TS theory may not always be applicable even to a well-known textbook organic reaction.

7.
J Org Chem ; 76(10): 3937-45, 2011 May 20.
Article in English | MEDLINE | ID: mdl-21486082

ABSTRACT

Measurements of rate constants and substituent effects for three important elementary steps of proton-transfer reactions of phenylnitromethane were reported. The Hammett ρ values for the deprotonation of ArCH(2)NO(2) with OH(-), protonation of ArCH═NO(2)(-) with H(2)O, and protonation of ArCH═NO(2)(-) with HCl were determined in aqueous MeOH at 25 °C. Comparison of these experimentally observed ρ values with those calculated at B3LYP/6-31G* revealed that aci-nitro species (ArCH═NO(2)H), which is formed on the O-protonation of ArCH═NO(2)(-), does not lie on the main route of the proton-transfer reaction. Analysis of the Brønsted plot implies that the proton-transfer reaction of most XC(6)H(4)CH(2)NO(2) exhibits nitroalkane anomaly, but not for p-NO(2)C(6)H(4)CH(2)NO(2), and that the transition state charge imbalance is an origin of anomaly.

8.
Chemistry ; 17(4): 1230-7, 2011 Jan 24.
Article in English | MEDLINE | ID: mdl-21243689

ABSTRACT

The mechanistic dichotomy between concerted E2 and stepwise E1cb of the base-promoted elimination of 2-aryl-3-chloro-2-R-propanols was examined computationally at the HF, M05-2X, and MP2 levels of theory. Optimizations of transition states (TSs) and reaction intermediates, and intrinsic reaction coordinates (IRC) calculations showed that there was a single reaction route for each substrate, and that the mechanism could be changed from E2 to E1cb by making a carbanion intermediate more stable through the introduction of electron-withdrawing substituents. Molecular dynamics simulations revealed that trajectories started at a single TS led directly to two product regions; the carbanion intermediate region in the E1cb mechanism, and the alkene product region in the E2 mechanism, through path bifurcation after the TS. The present system is a new example of bifurcation in reactions of closed-shell molecules. The overall reaction mechanism changes dynamically from E2 to E1cb by a gradual change in the ratio of E2 and E1cb trajectories, rather than a path switch in concurrent pathways.

10.
J Am Chem Soc ; 132(10): 3413-22, 2010 Mar 17.
Article in English | MEDLINE | ID: mdl-20166731

ABSTRACT

The N(2) liberation from iminodiazonium ion (2-X) is a key step of the Schmidt rearrangement of ketones. Molecular orbital calculations showed that two concurrent reaction channels, syn-benzyl fragmentation and anti-Me rearrangement, exist for syn-2, whereas anti-2-X proceeds via a single TS. Substituent effect analyses of the reactions of syn-2-X gave concave-upward plots, typical for a concurrent reaction mechanism. On the other hand, the reactions of anti-2-X gave linear Hammett plots, indicative of a single reaction mechanism for all anti-2-X. IRC calculations, however, revealed that the TS led to either an anti-benzyl rearrangement or an anti-benzyl fragmentation product depending on the substituent. Thus, the change of the mechanism (identity of the product) could not be detected by the Hammett plots. Ab initio dynamics simulations for anti-2-X were found to follow the IRC path for X = p-NO(2), giving the rearrangement product, and almost so for X = p-MeO, giving the fragmentation products. However, in borderline cases where X is less donating than p-MeO and less withdrawing than p-NO(2), the trajectories did not follow the minimum energy path on the potential energy surface but gave both rearrangement and fragmentation products directly from the single TS. This is a novel example of path bifurcation for a closed shell anionic reaction. It was concluded that a reactivity-selectivity argument based on the traditional TS theory might not always be applicable even to a well-known textbook organic reaction.

11.
J Org Chem ; 74(3): 1268-74, 2009 Feb 06.
Article in English | MEDLINE | ID: mdl-19123838

ABSTRACT

Proton-transfer reactions of two systems, ionization of a series of small carbon acids in water (the Pearson system) and reactions of substituted phenylnitromethanes, were examined in detail computationally. Comparison of experimental reactivity and pK(a) with calculated relative activation barrier and reaction energy for the Pearson system suggested that the origin of the well-know nitroalkane anomaly does not reside in the reactivity but in the equilibrium. For the reactions of substituted phenylnitromethanes, proton transfers among three species, PhCH(2)NO(2), PhCHNO(2)(-), and PhCH=NO(2)H, were examined, and the role of the aci-nitro species (PhCH=NO(2)H) was evaluated on the basis of its stability and reactivity. Protonation of PhCHNO(2)(-) by H(2)O was suggested to occur kinetically on the oxygen site, but due to its instability PhCH=NO(2)H does not contribute to the overall reaction energetics. The protonation of PhCHNO(2)(-) under acidic conditions occurs on the oxygen site to give PhCH=NO(2)H both kinetically and thermodynamically. The aci-nitro species thus formed appears to give PhCH(2)NO(2) via intramolecular H(2)O-mediated proton transfer, but a possibility of the route through PhCHNO(2)(-)-C-protonation would not be fully eliminated.

12.
J Comput Chem ; 30(3): 358-65, 2009 Feb.
Article in English | MEDLINE | ID: mdl-18613068

ABSTRACT

The gas-phase base-induced bimolecular elimination (E2) reactions at saturated carbon with 13 bases, B(-) + CH3CH2Cl --> BH + CH2=CH2 + Cl(-) (B = HO, CH3O, CH3CH2O, FCH2CH2O, ClCH2CH2O, Cl, Br, FO, ClO, BrO, HOO, HSO, and H2NO), were investigated with the high-level G2(+) theory. It was found that all alpha-bases with adjacent lone pair electrons examined exhibited downward deviations from the correlation line between the overall barriers and proton affinities for the normal bases without adjacent lone pair electrons, indicating the existence of the alpha-effect in the gas phase E2 reactions. The sizes of the alpha-effect for the E2 reaction, DeltaH(alpha)(E2), span a smaller range if the alpha-atoms are on the same column in the periodic table, in contrast to the corresponding S(N)2 reactions, where the DeltaH(alpha)(S(N)2) values significantly decrease from an upper to a lower column. The origin of the alpha-effects in E2 reactions can be interpreted by the favorable orbital interaction between the lone-pair electrons and positively charged anti-bonding orbital. It is worth noticing that the neighboring electron-rich pi lobe instead of lone pair electrons could also cause the alpha-effect in E2 reaction.


Subject(s)
Computer Simulation , Models, Chemical , Gases/chemistry
13.
J Comput Chem ; 30(1): 40-50, 2009 Jan 15.
Article in English | MEDLINE | ID: mdl-18504778

ABSTRACT

Fragment Molecular Orbital based-Molecular Dynamics (FMO-MD, Komeiji et al., Chem Phys Lett 2003, 372, 342) is an ab initio MD method suitable for large molecular systems. Here, FMO-MD was implemented to conduct full quantum simulations of chemical reactions in explicit solvation. Several FMO-MD simulations were performed for a sphere of water to find a suitable simulation protocol. It was found that annealing of the initial configuration by a classical MD brought the subsequent FMO-MD trajectory to faster stabilization, and also that use of bond constraint in the FMO-MD heating stage effectively reduced the computation time. Then, the blue moon ensemble method (Sprik and Ciccotti, J Chem Phys 1998, 109, 7737) was implemented and was tested by calculating free energy profiles of the Menschutkin reaction (H3N + CH3Cl --> +H3NCH3 + Cl-) in the presence and absence of the solvent water via FMO-MD. The obtained free energy profiles were consistent with the Hammond postulate in that stabilization of the product by the solvent, namely hydration of Cl-, shifted the transition state to the reactant-side. Based on these FMO-MD results, plans for further improvement of the method are discussed.


Subject(s)
Computer Simulation , Models, Chemical , Solvents/chemistry , Water/chemistry , Kinetics , Quantum Theory , Solubility , Thermodynamics
14.
J Org Chem ; 74(1): 188-96, 2009 Jan 02.
Article in English | MEDLINE | ID: mdl-19053588

ABSTRACT

The reversible deprotonation of 3(2H)-furanone (3H-O) and 3(2H)-thiophenone (3H-S) by a series of delocalized carbanions and by CN(-), and the identity proton transfer of 3H-O to its conjugate base (3(-)-O) and of 3H-S to 3(-)-S have been studied at the MP2//6-31+G** level. The main objective has been to examine to what extent the aromaticity of 3(-)-O and 3(-)-S is expressed at the transition state of these reactions and how the intrinsic barriers are affected by the transition state aromaticity. Aromaticity parameters such as NICS values, HOMA and Bird Indices indicate a disproportionately high degree of aromatic stabilization of the transition state. This stabilization results in a reduction of the intrinsic barriers which is most clearly manifested in the identity reactions. However, these reductions are relatively modest compared to those reported previously for the identity proton transfers from the benzenium ion to benzene and of cyclopentadiene to its conjugate base, reflecting the smaller aromatic stabilization of 3(-)-O and 3(-)-S compared to those of benzene and cyclopentadienyl anion.


Subject(s)
4-Butyrolactone/analogs & derivatives , 4-Butyrolactone/chemistry , Gases/chemistry , Protons , Quantum Theory , Thermodynamics , Thiophenes/chemistry
16.
Phys Chem Chem Phys ; 10(6): 844-9, 2008 Feb 14.
Article in English | MEDLINE | ID: mdl-18231687

ABSTRACT

Two-dimensional free-energy surfaces are calculated for alkyl chloride/chloride exchange/inversion reactions: Cl(-)+RCl (R=Me and t-Bu) surrounded by one hundred H(2)O molecules as a model of solvent. The methodology of free-energy calculation by perturbation theory based on a mixed-Hamiltonian model (QM/MM) combined with Monte Carlo sampling of the solvent configurations was used to obtain the changes in solvation free energy. We devised a special procedure to analyze the two-dimensional free-energy surfaces to gain unique insight into the differences in the reaction mechanisms between the two systems. The inversion reaction path for R=t-Bu on the free-energy surface is found to proceed in an asynchronous way within a concerted framework via the ion-pair region. This is in contrast to the R=Me system that proceeds as a typical S(N)2 reaction.


Subject(s)
Chlorides/chemistry , Computer Simulation , Hydrocarbons, Chlorinated/chemistry , Models, Chemical , Quantum Theory , Thermodynamics , Energy Transfer , Surface Properties , Water/chemistry
17.
J Org Chem ; 72(20): 7605-24, 2007 Sep 28.
Article in English | MEDLINE | ID: mdl-17760461

ABSTRACT

Reactions of isocyanates XNCO (e.g., X = p-An, Ph, i-Pr) with (MeO)2P(=O)CH2CO2R [R = Me, CF3CH2, (CF3)2CH] gave 15 formal "amides" (MeO)2P(=O)CH(CO2R)CONHX (6/7), and with (CF3CH2O)2P(=O)CH2CO2R [R = Me, CF3CH2] they gave eight analogous amide/enols 17/18. X-ray crystallography of two 6/7, R = (CF3)2CH systems revealed Z-enols of amides structures (MeO)2P(=O)C(CO2CH(CF3)2)=C(OH)NHX 7 where the OH is cis and hydrogen bonded to the O=P(OMe)2 group. The solid phosphonates with R = Me, CF3CH2 have the amide 6 structure. The structures in solution were investigated by 1H, 13C, 19F, and 31P NMR spectra. They depend strongly on the substituent R and the solvent and slightly on the N-substituent X. All systems displayed signals for the amide and the E- and Z-isomers. The low-field two delta(OH) and two delta(NH) values served as a probe for the stereochemistry of the enols. The lower field delta(OH) is not always that for the more abundant enol. The % enol, presented as K(enol), was determined by 1H, 19F, and 31P NMR spectra, increases according to the order for R, Me < CF3CH2 < (CF3)2CH, and decreases according to the order of solvents, CCl4 > CDCl3 approximately THF-d8 > CD3CN >DMSO-d6. In DMSO-d6, the product is mostly only the amide, but a few enols with fluorinated ester groups were observed. The Z-isomers are more stable for all the enols 7 with E/Z ratios of 0.31-0.75, 0.15-0.33, and 0.047-0.16 when R = Me, CF3CH2, and (CF3)2CH, respectively, and for compounds 18, R = Me, whereas the E-isomers are more stable than the Z-isomers. Comparison with systems where the O=P(OMe)2 is replaced by a CO2R shows mostly higher K(enol) values for the O=P(OMe)2-substituted systems. A linear correlation exists between delta(OH)[Z-enols] activated by two ester groups and delta(OH)[E-enols] activated by phosphonate and ester groups. Compounds (MeO)2P(=O)CH(CN)CONHX show

18.
J Org Chem ; 72(15): 5660-7, 2007 Jul 20.
Article in English | MEDLINE | ID: mdl-17590049

ABSTRACT

The origin of enhanced reactivity of alpha-nucleophiles in SN2 reactions was examined on the basis of computational results at the high level G2(+) method for 22 gas-phase reactions: Nu- + RCl --> RNu + Cl- [R = Et and i-Pr; Nu- = HO-, CH3O-, HS-, Cl-, Br-, NH2O-, HOO-, FO-, HSO-, ClO-, and BrO-]. The results clearly indicate the existence of the alpha-effect, whose size varies depending on the R group and the identity of the alpha-atom. The alpha-effect is larger for i-PrCl than EtCl, and for an alpha-nucleophile with a harder alpha-atom. Analyses of the present results, together with previously reported ones for MeF and MeCl reactions, reveal that several rationales so far presented to explain the alpha-effect, such as thermodynamic product stability, transition state (TS) tightness, electrostatic interaction, ET rationale, and polarizability, cannot explain the observed size of the alpha-effect. The importance of deformation energy on going from the reactant to the TS is presented.

19.
J Org Chem ; 72(14): 5297-312, 2007 Jul 06.
Article in English | MEDLINE | ID: mdl-17559283

ABSTRACT

Twenty open-chain mono-, di-, and trialkyl and aryl-N-substituted cyanomalonamides R2R1NCOCH(CN)CONHR3 were prepared. In solution, signals for both amide and a single enol are mostly observed, despite the potential for E and Z isomeric enols. The equilibrium (KEnol) values between the amides and the enols were determined in different solvents by NMR spectra. They decrease on increasing the polarity of the solvent in the order CDCl3 approximately C6D6>THF-d8>(CD3)2CO>CD3CN>DMF-d7>DMSO-d6. For the R1R2NCOCH(CN)CONHR3 system when R1=R2=H, Me or R1=H, R2=Me, KEnol for R3 follows the order: C6F5>Ph>or=An>or= i-Pr>or= t-Bu, and for R1, R2:H, H>Me, H>Me, Me in all solvents. A unique feature is the appreciable % enol in DMSO-d6 when R1=R2=H, in contrast with enol systems with other electron-withdrawing groups (EWGs). Calculations (B3LYP/6-31G**) corroborate the higher KEnol values for less alkyl-substituted systems, showing that in the most stable conformer when R1=H, R2=R3=Me the N-hydrogens are closer to the CN group. The order of promoting substituents for enol of amide formation is CONH2>CO2CH2CF3>CO2Me>CONHMe. The solid-state structures of the isolated species, determined by X-ray crystallography, were either amides or enols, and a higher KEnol(CDCl3) value does not ensure a solid enol structure. In no system were both solid species isolated. The X-ray structures of the enols were temperature-dependent. In most cases, the difference between the O-H and O...H bond lengths at low temperature were appreciable, but they become closer at the higher temperature. Similar tendency for either the C=C/C-C or the C-O/C=O bonds was observed. This is ascribed to a hydrogen shift between two regioisomeric enols in an asymmetric double-well potential, which becomes faster at a higher temperature. Calculations show that the enol structures are nonsymmetrical, resembling the lower temperature structures, even when they are chemically symmetrical, but the energy differences between the two regioisomers are <1 kcal. The hydrogen bonds in the enol moiety are strong, with O...O distances <2.45 A, and are resonance-assisted hydrogen bonds. IR spectra in solution and the solid state qualitatively corroborate the NMR and X-ray structure determination.

20.
Chemistry ; 13(2): 677-82, 2007.
Article in English | MEDLINE | ID: mdl-17009371

ABSTRACT

As a continuing theoretical study on the alpha-effect in the S(N)2 reactions at saturated carbon centers, 28 gas-phase reactions have been examined computationally by using the high-level G2(+) method. The reactions include: Nu(-)+CH(3)X-->CH(3)Nu+X(-) (X=F and Cl; Nu(-)=HO(-), HS(-), CH(3)O(-), Cl(-), Br(-), HOO(-), HSO(-), FO(-), ClO(-), BrO(-), NH(2)O(-), and HC(==O)OO(-)). It was found that all alpha-nucleophiles examined exhibit downward deviations from the correlation line between the overall barriers and proton affinities for normal nucleophiles, indicating the existence of the alpha-effect in the gas phase. The transition states (TS) for the alpha-nucleophiles are characterized by less advanced C--X bond cleavages than the normal nucleophiles, leading to smaller deformation energies and overall barriers. The size of the alpha-effect is related to the electron density on the alpha-atom, and increases when the position of alpha-atom is changed from left to right and from bottom to top in the periodic table. The reaction with CH(3)F exhibits a larger alpha-effect than that with CH(3)Cl, which can be explained by a later TS and a more positively charged methyl group at the TS for CH(3)F, [NuCH(3)F](- not equal). Thus, a higher electron density on the alpha-atom and a more positive methyl moiety at the TS result in a larger alpha-effect.

SELECTION OF CITATIONS
SEARCH DETAIL
...