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 31 毫秒
1.
Infrared multiphoton dissociation spectra of protonated p-aminobenzoic acid generated by electrospray ionization (ESI) from aqueous methanol and acetonitrile solutions were recorded in the gas phase from 2800-4000 cm(-1). The O-protonated ion is more stable than the N-protonated structure in the gas phase, whereas the opposite is true in both solutions. When CH(3)OH/H(2)O was used as the ESI solvent, only the O-protonated ion was observed. In contrast, a 70:30 mixture of the O- and N-protonated species were produced from CH(3)CN/H(2)O. These structural assignments are based on an assortment of experimental data (action spectra, photofragments, photofragmentation kinetics, and H/D exchange) and are fully supported by extensive computations. This work shows that ESI can lead to isomerization and that the ionization site may be varied by changing the solvent from which the substrate is analyzed.  相似文献   

2.
Protonation of typical unstrained amides and lactams is heavily favored at oxygen. In contrast, protonation of the highly distorted lactam 1-azabicyclo[2.2.2]octan-2-one is heavily favored at nitrogen. What structures occupy "crossover boundaries" where N- and O-protonation are nearly equienergetic? Density function theory calculations at the B3LYP/6-31G* level, as well as QCISD(T)/6-31G* calculations, predict that 1-azabicyclo[3.3.1]nonan-2-one favors N-protonation at nitrogen only very slightly (<2.0 kcal/mol; "gas phase") over O-protonation. (1)H and (13)C NMR as well as ultraviolet (UV) studies of this lactam, in its combination with sulfuric acid, confirm predominant protonation at nitrogen. Although the calculations very slightly favor the N-protonated chair-chair conformation, experimental spectra clearly support the N-protonated boat-chair. Broadened resonances in the (13)C NMR spectrum suggest an exchange phenomenon. Variable-temperature studies of the (13)C NMR spectra support dynamic exchange between the major tautomer (N-protonated) and the minor tautomer (O-protonated) in a roughly 4:1 mixture. The findings also support the published prediction that a twisted bridgehead lactam with the nitrogen lone pair (n(N)) as HOMO will protonate at nitrogen.  相似文献   

3.
The acid-promoted hydrolysis of 2,4,4-trimethyloxazolin-5-one (TMO) is studied employing the density functional theory (B3LYP) method in conjunction with the 6-31++G(d,p) basis set. Two types of reaction mechanism, N-protonated and O-protonated, are considered, involving protonation at the nitrogen and carbonyl oxygen of TMO to activate the C2 and C5 atoms, respectively, in favor of attack by water molecules. In the N-protonated pathway, the nucleophilic water molecule attacks the activated C2 atom, with a proton transfer from the water molecule to the oxygen atom attached to C2 and the fission of the C2-O bond, leading to a cis ring-opening product (N-acyl-alpha-amino isobutyric acid). While, in the O-protonated pathway, the nucleophilic water molecule attacks the activated carbonyl C5 atom, accompanied by a proton transfer from the water molecule toward the nitrogen atom of oxazole ring and the cleavage of C5-O bond; as a result, a corresponding trans product is generated. The water-assisted hydrolysis reactions are also examined together. A local microhydration model, in which an extra water molecule was added to obtain a continuous H-bond network around the reaction centers, was adopted to mimic the system for the two types of reaction processes. In addition, bulk solvent effect is introduced by use of the conductor-like polarizable continuum model (CPCM). Our computational results in kinetics and thermodynamics clearly manifest that the O-protonated pathway with the nucleophilic attack at the carbonyl C5 atom is more favorable than the N-protonated one, in nice agreement with the available experimental conclusion.  相似文献   

4.
[reaction: see text] We present an ab initio study of the acid-promoted hydrolysis reaction mechanism of N-formylaziridine in comparison with formamide. Since the rate of amide hydrolysis reactions depends on the formation of the tetrahedral intermediate, we focused our attention mainly on the reactant complex, the tetrahedral intermediate, and the transition state connecting these two stationary points. Geometries were optimized using the density functional theory, and the energetics were refined using ab initio theory including electron correlation. Solvent effects were investigated by using polarizable continuum method calculations. The proton-transfer reaction between the O-protonated and N-protonated amides was investigated. In acidic media, despite that the N-protonated species is more stable than the O-protonated one, it is predicted that both N-protonated and O-protonated pathways compete in the hydrolysis reaction of N-formylaziridine.  相似文献   

5.
[reaction: see text] Data on the apparent dipole moment of thianthrene-5-oxide (1) and (1)H NMR spectra in different solvents support the conformational mobility of 1, which flaps between two limit boat conformations with the sulfinyl group in pseudoequatorial and pseudoaxial positions, respectively. The conformational equilibrium of 1 occurs too fast for the (1)H NMR (500 MHz) time-scale even at -130 degrees C, and the equilibrium constant has not been determined. The apparent dipole moments of 1 in n-hexane and 1,4-dioxane and the (1)H NMR spectra of 1 and the model compounds cis- and trans-thianthrene-5,10-dioxides (2) and thianthrene (5) in different solvents and at various temperatures confirm that the relative position of the conformational equilibrium of 1 is solvent-dependent, and more polar solvents favor the conformation with the sulfoxide group in the pseudoaxial position (1(')(ax)). Variable-temperature (1)H NMR spectra have established the interconversion barrier of trans-2 and confirmed that the conformational equilibrium of cis-2 is strongly displaced toward the conformation with both sulfinyl groups in the pseudoequatorial position. The (1)H NMR data support the transannular interaction of the functional groups in 1 and trans-2.  相似文献   

6.
The rates of the reactions of benzhydrylium ions (diarylcarbenium ions) with solvent mixtures of variable composition (water/acetonitrile, methanol/acetonitrile, ethanol/acetonitrile, ethanol/water, and trifluoroethanol/water) have been determined photometrically by conventional UV-vis spectroscopy, stopped-flow methods, and laser flash techniques. It has been shown that the first-order rate constants follow the previously published relationship log k(20 degrees C) = s(N + E), where E is an empirical electrophilicity parameter, N is an empirical nucleophilicity parameter, and s is a nucleophile-specific slope parameter. From plots of log k versus E of the benzhydrylium ions are derived the solvent nucleophilicity parameters s and N, the latter of which are designated as N1 to emphasize that their use in the quoted correlation equation gives rise to first-order rate constants. A linear correlation between N1 and Kevill's solvent nucleophilicity NT based on S-methyldibenzothiophenium ions is reported, which allows one to interconvert the two sets of data. Because the N1 values are directly comparable to the previously reported nucleophilicity parameters N for pi-systems (www.cup.uni-muenchen.de/oc/mayr/), the systematic design of Friedel-Crafts reactions with solvolytically generated carbocations becomes possible.  相似文献   

7.
Variable-temperature NMR experiments and ab initio density functional calculations were carried out to investigate the conformation interconversion of novel chiral 3-alkyl-3,4-dihydro-2H-benzo[1,4]oxazine derivatives. With CDCl3 as the solvent, the coalescence temperatures of H2, H3, H11, and H19 of product 1 are about 289, 304, 292, and 316 K, with the corresponding activation free energies at 58.0 +/- 6.7, 60.9 +/- 7.1, 58.3 +/- 6.8, and 59.6 +/- 6.9 kJ.mol(-1), respectively. When dimethyl sulfoxide (DMSO-d6) was used as the solvent, 1H and 13C NMR signals were completely assigned at 375 K. The effects of solvent and temperature were investigated through a polarizable continuum model. At each theoretical level (MP2 or B3LYP), the changing tendencies of the calculated activation free energies and interconversion rates agree well with those of the NMR results. In addition, the interconversion rate at each specified temperature was calculated to be about 1.5 times faster in DMSO-d6 than in CDCl3. Accordingly, we failed to observe the coalescence phenomena of H3 and H19 in DMSO-d6 by NMR measurements from 296 to 375 K. The substitution effect at the R1-R5 positions was considered using density functional calculations, with the activation barriers decreasing as follows: product 6 > 3 > 1 > 7 > 2. This sequence is consistent with that of the reaction heats, except for product 7, implying that the interconversion processes may be thermodynamically controlled. Surprisingly, the substituted groups near the acetyl group in product 2 and 7 do not elevate the activation barrier but, instead, lower it somewhat, with the possible reasons for this provided in the paper.  相似文献   

8.
[reaction: see text] The solution and chelation properties of 2-thienyllithium reagents with potential amine and ether chelating groups in the 3-position and related model systems have been investigated using low temperature 6Li, 7Li, 13C, and 31P NMR spectroscopy, 15N-labeling, and the effect of solvent additives. In THF-ether mixtures at low temperature 3-(N,N-dimethylaminomethyl)-2-thienyllithium (4) is ca. 99% dimer (which is chelated) and 1% monomer (unchelated), whereas 3-(methoxymethyl)-2-thienyllithium (5) is <10% dimer. Compound 5 crystallizes as a THF-solvated dimer, but there is no indication that the ether side chain is chelated in solution. Both 4 and 5 form PMDTA-complexed monomers almost stoichiometrically, similar to the model compound 2, in sharp contrast to phenyl analogues, which show very different behavior. The barriers to dimer interconversion are ca. 2 kcal/mol lower and chelation is significantly weaker in the 2-thienyllithium reagents than in their phenyl analogues.  相似文献   

9.
The mechanism of the A2 hydrolysis of acetamide has been investigated theoretically using MNDO Method. Fully optimized geometries of all species at the stationary points corresponding to energy minima and energy maxima along the reaction coordinate are determined for the two reaction paths: the rate-determining nucleophilic attack of water on the carbonyl carbon (i) of the O-protonated tautomer and (ii) of the N-protonated form. Results show that the latter provides a lower energy path by 7.5 Kcal/Mol compared to the former.Tetrahedral species' found were not at the energy minima but at or near the saddle points. Optimizied structures and formal charges on heavy atoms showed that the bond interchange with the concurrent proton interchange takes place at the rate-determining step. The negative charge on N atom was found to increase in the rate-determining step relative to that of the ground state, the O-protonated acetamide, and hence substitution of electron withdrowing group on N is predicted to depress the activation energy in agreement with the experimental results.  相似文献   

10.
The closo-[Sn9M(CO)3]4-ions where M = Cr (1), Mo (2), W (3) were prepared from [LM(CO)3] precursors (L=mesitylene, cycloheptatriene), K4Sn9. and 2,2,2-cryptand in ethylenediamine/toluene solvent mixtures. The [K(2,2,2-cryptand)]+ salts are very air and moisture sensitive and have been characterized by IR, 119Sn, and 13C NMR spectroscopy and single-crystal X-ray diffraction studies. Complexes 1-3 form bicapped square-antiprismatic 10-vertex 22-electron closo structures in which the [M(CO)3] units occupy cluster vertices. For 1 and 2, the clusters have C4. symmetry in the solid state in which the [M(CO)A] fragments occupy capping positions with Sn9(4-) ions that are bound to the metal in an 4 fashion. For 3, the [M(CO)3] fragment occupies a position in the square plane with an eta/5-Sn9(4-) ion and C(s) point symmetry. For 1-3, a dynamic equilibrium exists between the eta4 and eta5 structures yielding three 119Sn NMR signals that reflect the three chemically distinct Sn environments of the higher symmetry C(4v) structure. The 119Sn NMR chemical shifts and coupling constants show solvent and temperature dependencies due to the equilibrium process. A triangular-face rotation mechanism is proposed to describe the dynamic behavior.  相似文献   

11.
The dissolution and complex formation of fluoroaluminates in two eutectic alkalifluoride mixtures, NaF-KF (FNAK) and LiF-NaF-KF (FLINAK), have been investigated by Raman, NMR, and thermal analysis. Melting and dissolution took place stepwise. The eutectic alkalifluoride mixtures with minor amounts of dissolved fluoroaluminate salts started melting at around 460 and 740 degrees C for FLINAK and FNAK mixtures, respectively. Total melting/dissolution of mixtures with 9-11 mol % aluminum fluoro salts added took place near 780 degrees C in the FLINAK solvent and at approximately 900 degrees C for FNAK solutions. The solidified melts were characterized by Raman bands at 561 (nu(1)), 391 (nu(2)), and 328 cm(-1) (nu(5)) and a (27)Al NMR chemical shift near 0 ppm originating from isolated AlF(6)(3-) octahedral ions. The Raman and NMR signals due to AlF(6)(3-) were also observed at temperatures where the samples were only partly melted. Upon total melting, a pronounced dissociation of AlF(6)(3-) into AlF(5)(2-) and fluoride ions took place. At even higher temperatures, the equilibrium was displaced in favor of AlF(5)(2-) in the FNAK solvent. The AlF(5)(2-) ion was characterized by an intensive Raman band at 558 cm(-1) and an increasingly positive (27)Al chemical shift with raising temperature, e.g., of 16 ppm at 935 degrees C.  相似文献   

12.
The thermal and photochemical solvolysis of the two stereoisomeric 2-phenyl-1-propenyl(phenyl)iodonium tetrafluoroborates has been investigated in alcoholic solvents of varying nucleophilicity. The product profiles and rates of product formation in the thermal reaction are all compatible with a mechanism involving cleavage of the vinylic C-I bond assisted by the group in the trans position (methyl or phenyl), always leading to rearranged products. Depending on the nucleophilicity of the solvent, the primarily formed cations may or may not further rearrange to more stable isomers. The less reactive Z compound also yields some unrearranged vinyl ether product in the more nucleophilic solvents via an in-plane S(N)2 mechanism. The mechanism of the photolysis involves direct, unassisted cleavage of the vinylic, and aromatic, C-I bond in an S(N)1 mechanism. This produces a primary vinyl cation, which is partially trapped prior to rearrangement in methanol. The unrearranged vinyl ethers are mainly formed with retention of configuration via a lambda3-iodonium/solvent complex in an S(N)i mechanism. Thermal and photochemical solvolyses of iodonium salts are complementary techniques for the generation of different cation intermediates from the same substrate.  相似文献   

13.
利用L-N, N-双(β-羟乙基)丝氨酸及L-N, N-双(β-羟乙基)苏氨酸与三乙氧基硅烷或氯烷基三乙氧基硅烷反应合成了具有手性的(4S)-4-羧基杂氮硅三环(1-5), 并运用IR, 1H NMR, EI-MS等手段表征了结构。证据显示存在着贯穿笼状结构的N→Si配键。  相似文献   

14.
Spectra of the N‐phenyl‐5,6‐dihydro‐2,4‐diphenylbenzo[h]quinolinium tetrafluoroborate (1) and of the N‐phenyl‐5,6,8,9‐tetrahydro‐7‐phenyldibenzo[c,h]acridinium tetrafluoroborate (2) were recorded in various solvents and temperatures. The analysis of the 1H‐NMR spectra of the tetrafluoroborate salt 1, recorded in acetone, acetonitrile, 1,1,2,2‐tetrachloroethane and chloroform, revealed the existence of an equilibrium between two conformers in solution. Tight ion‐pairing in chloroform led to a smaller barrier for interconversion between the two conformers. In more polar solvents, where the dihydrobenzoquinolinium exists as a free cation, theoretical calculations predicted larger barriers. The spectra of 1 in 1,1,2,2‐tetrachloroethane also varied with temperature, resembling at higher temperatures the spectrum in CDCl3 and at 300K spectra in more polar media. Spectra of 2 did not vary with the solvent or the temperature, in an indication of a much higher barrier to conformational interconversion, because of a greater steric hindrance between the N‐phenyl substituent and the dihydrobenzo rings. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

15.
Secondary squaramides have considerable potential as hydrogen bond donors and acceptors. In CHCl(3) both, anti- and syn-squaramide rotamers are observed by NMR. The energetic barrier for anti/syn mutual interconversion determined by complete band shape analysis is approximately 63 kJ mol(-1). As in proline derivatives, a low rotational barrier is crucial for the design of foldable modules. In this paper, folding based on the low rotational barrier of squaramides is driven by donor atoms (N or O) located in the gamma position of an alkyl chain of a secondary squaramide. We demonstrate that the resulting minimal module exists as a folded conformer through the formation of a nine-membered ring stabilized by intramolecular hydrogen bonding. Molecular mechanics calculations and NMR studies support the existence of these folded conformers. The intramolecularly hydrogen bonded conformers are clearly visible even in CHCl(3)-EtOH mixtures. Folding occurs even in pure ethanol. As an indirect test, we studied the effectiveness of macrocyclization reactions in pure ethanol that require an effective templating effect to take place. The high yields obtained support the dominant role of a folded conformer even in this solvent.  相似文献   

16.
The photochemically generated 4-(N,N-dimethylamino)phenyl cation adds to norbornene giving a phenylnortricyclene and various 2-exo substituted phenylnorbornanes (main isomers, 3-endo-, 7-anti- and 7-syn-phenyl). Acetamides are obtained in MeCN and ethers in alcohols (MeOH, iso-PrOH, tert-BuOH, CF3CH2OH). The product distribution is closely reminiscent of that obtained in the solvolysis of 2-norbornyl derivatives, supporting that the reaction offers a novel access to a 'non classical' 2-norbornyl cation. The fate of this cation is determined by the basicity/nucleophilicity of the solvent.  相似文献   

17.
Values for the specific rates of solvolysis of the benzhydryldimethylsulfonium ion in 34 solvents have been analyzed using various forms of the extended Grunwald-Winstein equation. The specific rates are insensitive toward changes in solvent nucleophilicity (N(T)) values, and they correlate best against a combination of Y(+) values (based on the solvolyses of the 1-adamantyldimethylsulfonium ion) and aromatic ring parameter (I) values. Common-molecule return is observed, being especially powerful in solvents rich in fluoro alcohol; the logarithm of the associated mass law constant correlates inversely with the solvent N(T) values. The product selectivities in ethanol-water mixtures are also consistent with an S(N)1 mechanism for the solvolyses.  相似文献   

18.
Two classes of diastereomerically enriched chiral tridentate ligands incorporating either two oxazolidine and one pyridine ( 1) or two pyridine and one oxazolidine ( 2a-c) donor groups have been made in a high-yielding modular fashion from readily available enantiopure amino alcohols and aldehydes. Both ligand classes readily formed metal complexes via 1:1 reactions with trans-PdCl 2(PhCN) 2. The compounds Pd( 1)Cl 2 and Pd( 2a-c)Cl 2 were formed as mixtures of 3 and 2 diastereomers, respectively, owing to indeterminate absolute configuration at the C (2) position of their constituent oxazolidine rings. Within each diastereomeric manifold, the metal complexes existed as equilibrium mixtures of bi- and tridentate isomers in solution, the interconversion between which was very rapid even at -50 degrees C. The fluxional nature of the compounds was inferred from a combination of (1)H and (15)N NMR spectroscopic and solution conductivity data. Substitution of one chloride ligand for hexafluorophosphate gave as mixtures of diastereomers the salts [Pd( 1-kappa N,kappa (2) N')Cl]PF 6 ( 8) and [Pd( 2a-c-kappa N,kappa N',kappa N')Cl]PF 6 ( 12a-c) in which the ligands were coordinated through all three N-donors. A single recrystallization of 8 gave in optically pure form the major diastereomer 8 ( maj ), which was characterized crystallographically. Complexes of 2a-c differed substantially from those of the bis(pyridylmethyl)amine (bpma) ligand family with which they shared direct atom-for-atom connectivity in the coordinating groups. The amines are known to form exclusively the static tridentate complexes [PdCl(bpma-kappa N,kappa (2) N')]Cl; the difference was attributed to torsional strain associated with the rigid oxazolidine ring in tricoordinated 2a-c.  相似文献   

19.
Rate constants for the reactions of Laser flash photolytically generated benzhydrylium ions (diarylcarbenium ions) with halide ions have been determined in various solvents, including neat and aqueous acetonitrile as well as some alcohols. Substitution of the rate constants into the correlation equation log k = s(N + E) yields the nucleophilicity parameters N for the halide ions in different solvents. Linear correlations with negative slopes are found between the nucleophilicity parameters N for Cl(-) and Br(-) in different solvents and the solvent ionizing powers Y of the corresponding solvents. Increasing halide solvation reduces the rates of carbocation/chloride combinations by approximately half as much as it increases the rates of ionizations of benzhydryl chlorides. Comparison of the solvent dependent nucleophilicity parameters N of halide anions and the nucleophilicity parameters N(1) for solvents yields a quantitative prediction of common ion rate depression, as demonstrated by the analysis of a variety of literature reported mass-law constants alpha. Combination of the rate constants for the reactions of benzhydrylium ions with halide ions (k(-)()(1)) reported in this work with the ionization constants of benzhydryl halides (k(1)) and the recently reported rate constants for the reactions of benzhydrylium ions with solvents (k(2)) yields complete quantitative free energy profiles for solvolysis reactions. The applicability of Hammond's postulate for interpreting solvolysis reactions can thus be examined quantitatively.  相似文献   

20.
The conformational equilibrium of 1-alkyl-3-hydroxypiperidines in n-heptane and aqueous solution is investigated. Several methods, e.g. IR and NMR spectroscopy, are discussed. The conformation trapping in D2SO4-D2O mixtures, and subsequent analysis of the 1H and 13C NMR spectra, is proven to be the most reliable method. In an apolar solvent the conformation with an axial hydroxyl group is always dominant (69%). In aqueous solution the same conformation predominates in the protonated form (56%, pH<5), while only 43% of this conformation is present in the free base (pH~11). A detailed 13C NMR study is described of 1-isopropyl-3-hydroxypiperidine in aqueous solution at pH between 2 and 11.  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