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.
With a view to clarifying analogies and differences between the mechanisms involved in the nitrosation of amino acids and secondary amines, we studied the kinetics of the nitrosation of five imino acids (azetidine-2-carboxylic acid, pyrrolidine-2-carboxylic acid, piperidine-2-carboxylic acid, piperidine-3-carboxylic acid, and piperidine-4-carboxylic acid) and of the ethyl esters of three of them. Reaction kinetics were determined by the initial rate method, by spectrophotometric monitoring of the concentration of nitroso amino acid formed. The presence of the ? COO? group in the amino acids opens a new mechanistic route for the nitrosation of the secondary amino group: a nitrosyl carboxylate formed initially acts as an internal nitrosating agent, resulting in intramolecular migration of ? N ? O from the carboxylate group to the secondary amino group. The observed order of the α?, β?, and γ-amino acids as regards the ease of N-nitrosation by this route is explained in terms of the relative energies of (a) the equatorial and axial orientations of the Cring? Ccarboxyl bond, and (b) the chair and boat forms of the piperidine ring. © 1994 John Wiley & Sons, Inc.  相似文献   

2.
Nitrite has long been considered a potential pre‐carcinogen for gastric cancer. Acidification of salivary nitrite, derived from dietary nitrate, produces nitrosative species such as NOSCN, NO+ and N2O3, which can form potentially carcinogenic N‐nitroso compounds. Ascorbic acid inhibits nitrosation by converting the nitrosative species into nitric oxide (NO). However, NO diffuses rapidly to adjacent lipids, where it reacts with oxygen to reform nitrosative species. Nitrosation has been studied in vitro in aqueous systems and less frequently in organic systems; however, there is a need to investigate acid‐catalysed nitrosation in a system combining aqueous and lipid environments, hence providing a physiologically relevant model. Here, we describe a two‐phase system, which can be used as a tool to understand acid‐catalysed nitrosation. Using gas chromatography/ion trap tandem mass spectrometry, we investigated the nitrosation of secondary amines as a function of the lipid phase composition and reaction mixing. An increased interface surface area was a driver for nitrosation, while incorporation of unsaturated fatty acids affected morpholine and piperidine nitrosation differently. Linoleic acid methyl esters did not affect morpholine nitrosation and only had a limited effect on N‐nitrosopiperidine formation, while incorporation of free linoleic acid to the lipid phase significantly reduced N‐nitrosopiperidine formation, but increased N‐nitrosomorpholine formation at low levels. The mechanisms driving these effects are thought to involve amine partitioning, polarity and unsaturated fatty acids acting as scavengers of nitrosating species, findings relevant to the nitrosative chemistry occurring in the stomach, where the gastric acid meets a range of dietary fats which are emulsified during digestion. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

3.
The fragmentation–rearrangement of peptide backbones mediated by nitrogen dioxide, NO2., was explored using di‐, tri‐, and tetrapeptides 8 – 18 as model systems. The reaction, which is initiated through nonradical N‐nitrosation of the peptide bond, shortens the peptide chain by the expulsion of one amino acid moiety with simultaneous fusion of the remaining molecular termini through formation of a new peptide bond. The relative rate of the fragmentation–rearrangement depends on the nature of the amino acids and decreases with increasing steric bulk at the α carbon in the order Gly>Ala>Val. Peptides that possessed consecutive aromatic side chains only gave products that resulted from nitrosation of the sterically less congested N‐terminal amide. Such backbone fragmentation–rearrangement occurs under physiologically relevant conditions and could be an important reaction pathway for peptides, in which sections without readily oxidizable side chains are exposed to the air pollutant NO2.. In addition to NO2.‐induced radical oxidation processes, this outcome shows that ionic reaction pathways, in particular nitrosation, should be factored in when assessing NO2. reactivity in biological systems.  相似文献   

4.
The kinetics of nitrosation of Methylurea (MU) in aqueous perchloric solution has been studied using two techniques: a dynamic spectrophotometric and a stopped-flow technique. The rate law obtained, whenpH was varied in the range 0.27–3.22, is $$---d[nitrite]/dt = f[MU][nitrite][H^ + ]/(1 + g/[H^ + ])$$ where [MU] and [nitrite] represent stoichiometric concentrations. At 288.0 K and μ=1.0M,f=(15.6±0.5)M ?2s?1 andg=(1.06±0.08) 10?3 M. This rate law becomes independent of the acidity of the solution when this is increased ([ClO4H])>1.00M). These results together with the activation of the nitrosation rate by the ionic strength and the negative value of the activation entropy shown that only the NO2H2 + or NO+ is the effective carrier for the nitrosation. Comparisons with the nitrosation of dimethylamine were also made leading to the conclusion that there is no simple explanation for the fact that the nitrosation via NO2H2 +/NO+ disappears when the nitrosable compound is of increased basicity.  相似文献   

5.
The kinetics of the reactions between sodium nitrite and phenol or m-, o-, or p-cresol in potassium hydrogen phthalate buffers of pH 2.5–5.7 were determined by integration of the monitored absorbance of the C-nitroso reaction products. At pH > 3, the dominant reaction was C-nitrosation through a mechanism that appears to consist of a diffusion-controlled attack on the nitrosatable substrate by NO+/NO2H2+ ions followed by a slow proton transfer step; the latter step is supported by the observation of basic catalysis by the buffer which does not form alternative nitrosating agents as nitrosyl compounds. The catalytic coefficients of both anionic forms of the buffer have been determined. The observed order of substrate reactivities (o-cresol ≈ m-cresol > phenol ≫ p-cresol) is explained by the hyperconjugative effect of the methyl group in o- and m-cresol, and by its blocking the para position in p-cresol. Analysis of a plot of ΔH# against ΔS# shows that the reaction with p-cresol differs from those with o- and m-cresol as regards the formation and decomposition of the transition state. The genotoxicity of nitrosatable phenols is compared with their reactivity with NO+/NO2H2+. © 1997 John Wiley & Sons, Inc.  相似文献   

6.
DFT investigations are carried out to explore the effective catalyst forms of DBU and H2O and the mechanism for the formation of 2,3‐dihydropyrido[2,3‐d]‐pyrimidin‐4(1H)‐ones. Three main pathways are disclosed under unassisted, water‐catalyzed, DBU and water cocatalyzed conditions, which involves concerted nucleophilic addition and H‐transfer, concerted intramolecular cyclization and H‐transfer, and Dimroth rearrangement to form the product. The results indicated that the DBU and water cocatalyzed pathway is the most favored one as compared to the rest two pathways. The water donates one H to DBU and accepts H from 2‐amino‐nicotinonitrile ( 1 ), forming [DBU‐H]+‐H2O as effective catalyst form in the proton migration transition state rather than [DBU‐H]+‐OH?. The hydrogen bond between [DBU‐H]+···H2O··· 1 ? decreases the activation barrier of the rate‐determining step. Our calculated results open a new insight for the green catalyst model of DBU‐H2O. © 2015 Wiley Periodicals, Inc.  相似文献   

7.
This contribution develops a comprehensive kinetic model of the N-nitrosation reaction mechanism, consisting almost entirely of elementary reaction steps and applies it to study the nitrosation of ammonia. The reaction mechanism features 26 species and 22 reactions, with 8 parallel reaction pathways for ammonia nitrosation and a side pathway for nitrous acid decomposition. We compiled forward and reverse rate constants for each of the reactions, either from the literature sources or by correlation with known rate and equilibrium constants. The concentration of each reaction species with respect to time can be obtained for any set of initial concentrations by invoking a simultaneous solution to the system of ordinary differential equations describing the reaction mechanism. The model successfully predicts previous experimental results for ammonia nitrosation, with and without the addition of the catalyst thiocyanate. The effect of pH on the rate and mechanism of ammonia nitrosation was studied with the model. For uncatalyzed nitrosation, the results indicate that between pH 6.0 and 1.5 the reaction proceeds predominantly via reaction with N2O3, whereas ON+ nitrosation becomes the preferred pathway below pH 1.5. ONSCN is the dominant nitrosating agent across the entire pH range studied when the nucleophile thiocyanate was added in appreciable quantities. However, with the weak nucleophile Cl, nitrosation by N2O3 and ON+ governed the reaction kinetics at high and low pH, respectively. We demonstrate that nitrosating agent formation is rapid and does not limit the rate of ammonia nitrosation; however, nitrosating agent formation could become the rate-limiting step for the nitrosation of highly reactive substrates. © 2007 Wiley Periodicals, Inc. Int J Chem Kinet 39: 645–656, 2007  相似文献   

8.
The recently synthesized ammonium dinitramide (ADN) is an ionic compound containing the ammonium ion and a new oxide of nitrogen, the dinitramide anion (O2N? N? NO2?). ADN has been investigated using high-energy xenon atoms to sputter ions directly from the surface of the neat crystalline solid. Tandem mass spectrometric techniques were used to study dissociation pathways and products of the sputtered ions. Among the sputtered ionic products were NH4+, NO+, NO2?, N2O2?, N2O, N3O4? and an unexpected high abundance of NO3?. Tandem mass spectra of the dinitramide anion reveal the uncommon situation where a product ion (NO3?) is formed in high relative abundance from metastable parent ions but is formed in very low relative abundance from collisionally activated parent ions. It is proposed that the nitrate anion is formed in the gas phase by a rate-determining isomerization of the dinitramide anion that proceeds through a four-centered transition state. The formation of the strong gas-phase acid, dinitraminic acid (HN3O4), the conjugate acid of the dinitramide anion, was observed to occur by dissociation of protonated ADN and by dissociation of ADN aggregate ions with the general formula [NH4(N(NO2)2)n] NH4+, where n = 1–30.  相似文献   

9.
Yanlong Kang 《Tetrahedron》2004,60(49):11219-11225
The use of simple calix[4]arenes 1a,b for NO2/N2O4 sensing and conversion is demonstrated, both in solution and in the solid state. Upon reacting with these gases, compounds 1a,b encapsulate reactive NO+ cations within their cavities with the formation of deeply colored (λmax∼570 nm) charge-transfer complexes 2a,b. Further functionalization of the calix[4]arene platform is reported for attachment to solid supports. Polymer-supported calixarene material 3 was prepared, which reversibly traps NO2/N2O4 with the formation of nitrosonium storing polymer 4. Material 4 was effectively used for nitrosation of amides.  相似文献   

10.
The mechanism of the pyrolysis reaction of carpronium chloride [(CH3)3N+? (CH2)3? COOCH3CI?] leading to γ-butyrolactone and tetramethylammonium chloride was investigated by means of thermal analysis, pyrolysis gas chromatography mass spectrometry and field desorption mass spectrometry, using deuterium labelling. The results indicated that carpronium chloride pyrolysed to yield equimolar amounts of γ-butyrolactone and tetramethylammonium chloride, methyl transfer occurred between N and O during the pyrolysis process. The mechanism is discussed on the basis of the experimental results, and with the aid of the theoretical results calculated by the CNDO/2 method. The mechanism presented is as follows. γ-Butyrolactone is formed by the intramolecular migration of the π-orbital of C?O to the carbon adjacent to [(CH3)3N]+ via a 5-membered ring transition state, accompanied by a bimolecular reaction between [(CH3)3N]+ and the CH3 of O? CH3, resulting in the formation of tetramethylammonium chloride in an amount equimolar with γ-butyrolactone.  相似文献   

11.
Reinvestigation of the Ar 1048-1067 Å photoelectron spectrum of NO2 using a completely NO-free sample suggests an upper limit for the adiabatic first ionization energy of this molecule consistent with the 9.75 eV photoionization value. The very low cross section process observed below 10.0 eV is interpreted by a mechanism involving autoionization followed by radiationless transition toward NO2+ ionic ground state.  相似文献   

12.
We report a novel 1:1 cocrystal of β‐alanine with dl ‐tartaric acid, C3H7NO2·C4H6O6, (II), and three new molecular salts of dl ‐tartaric acid with β‐alanine {3‐azaniumylpropanoic acid–3‐azaniumylpropanoate dl ‐tartaric acid–dl ‐tartrate, [H(C3H7NO2)2]+·[H(C4H5O6)2], (III)}, γ‐aminobutyric acid [3‐carboxypropanaminium dl ‐tartrate, C4H10NO2+·C4H5O6, (IV)] and dl ‐α‐aminobutyric acid {dl ‐2‐azaniumylbutanoic acid–dl ‐2‐azaniumylbutanoate dl ‐tartaric acid–dl ‐tartrate, [H(C4H9NO2)2]+·[H(C4H5O6)2], (V)}. The crystal structures of binary crystals of dl ‐tartaric acid with glycine, (I), β‐alanine, (II) and (III), GABA, (IV), and dl ‐AABA, (V), have similar molecular packing and crystallographic motifs. The shortest amino acid (i.e. glycine) forms a cocrystal, (I), with dl ‐tartaric acid, whereas the larger amino acids form molecular salts, viz. (IV) and (V). β‐Alanine is the only amino acid capable of forming both a cocrystal [i.e. (II)] and a molecular salt [i.e. (III)] with dl ‐tartaric acid. The cocrystals of glycine and β‐alanine with dl ‐tartaric acid, i.e. (I) and (II), respectively, contain chains of amino acid zwitterions, similar to the structure of pure glycine. In the structures of the molecular salts of amino acids, the amino acid cations form isolated dimers [of β‐alanine in (III), GABA in (IV) and dl ‐AABA in (V)], which are linked by strong O—H…O hydrogen bonds. Moreover, the three crystal structures comprise different types of dimeric cations, i.e. (AA)+ in (III) and (V), and A+A+ in (IV). Molecular salts (IV) and (V) are the first examples of molecular salts of GABA and dl ‐AABA that contain dimers of amino acid cations. The geometry of each investigated amino acid (except dl ‐AABA) correlates with the melting point of its mixed crystal.  相似文献   

13.
Crystals of maleates of three amino acids with hydrophobic side chains [L‐leucenium hydrogen maleate, C6H14NO2+·C4H3O4, (I), L‐isoleucenium hydrogen maleate hemihydrate, C6H14NO2+·C4H3O4·0.5H2O, (II), and L‐norvalinium hydrogen maleate–L‐norvaline (1/1), C5H11NO2+·C4H3O4·C5H12NO2, (III)], were obtained. The new structures contain C22(12) chains, or variants thereof, that are a common feature in the crystal structures of amino acid maleates. The L‐leucenium salt is remarkable due to a large number of symmetrically non‐equivalent units (Z′ = 3). The L‐isoleucenium salt is a hydrate despite the fact that L‐isoleucine is a nonpolar hydrophobic amino acid (previously known amino acid maleates formed hydrates only with lysine and histidine, which are polar and hydrophilic). The L‐norvalinium salt provides the first example where the dimeric cation L‐Nva...L‐NvaH+ was observed. All three compounds have layered noncentrosymmetric structures. Preliminary tests have shown the presence of the second harmonic generation (SGH) effect for all three compounds.  相似文献   

14.
Studies of nitrosation of norbornene and norbornadiene derivatives and dimethyl tricyclo[4.2.2.02,5]deca-3,7-diene-9,10-cis-endo-dicarboxylate demonstrated that nitrosation of alkenes with EtONO-PHal3, EtONO-POHal3 (Hal = Cl or Br), and EtONO-SOCl2 systems can afford nitroso halides in high yields without the formation of by-products (ketones and oximes). The reactions with 5-substituted norbornenes are nonregioselective. The trans dimer of endo-5-trifluoromethyl-cis-exo-2-chloro-3-nitrosobicyclo[2.2.1]heptane was studied by X-ray diffraction. Dedicated to Academician N. S. Zefirov on the occasion of his 70th birthday. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 9, pp. 2070–2080, September, 2005.  相似文献   

15.
Gabapentin [or 1‐(aminomethyl)cyclohexaneacetic acid], C9H17NO2, exists as a zwitterion [1‐(ammoniomethyl)cyclohexaneacetate] in the solid state. The crystal structures and bonding networks of two new monoclinic polymorphs (β‐gabapentin and γ‐gabapentin) are studied and compared with a previously reported gabapentin polymorph [α‐gabapentin: Ibers (2001). Acta Cryst. C 57 , 641–643]. All three polymorphs have extensive networks of hydrogen bonds between the NH3+ and COO groups of neighbouring molecules. In β‐gabapentin, there is an additional weak intramolecular hydrogen bond.  相似文献   

16.
Photoexcited (UO2NO3+)aq is shown to have different structure and thermodynamic formation functions from ground-state (UO2NO3+)aq and different photophysical and photochemical parameters from UO2+2(aq). Results are discussed with relation to the electron-transfer nature of the lowest excited state of UO2+2.  相似文献   

17.
Traditional catalytic procedures for oxidation of phenol produce environmentally undesirable wastes. As a consequence, there is a clear demand for development of an environmentally benign catalytic route for the selective oxidation of phenol. A series of zeolite-Y enslaved Mn(III) complexes with Schiff bases derived from vanillin furoic-2-carboxylic acid hydrazone (VFCH), vanillin thiophene-2-carboxylic acid hydrazone (VTCH), ethylvanillin thiophene-2-carboxylic acid hydrazone (EVTCH), and/or ethylvanillin furoic-2-carboxylic acid hydrazone have been synthesized and characterized by physico-chemical techniques. Catalytic oxidations of phenol using 30% H2O2 as an oxidant over [Mn(VTCH)2·2H2O]+-Y, [Mn(VFCH)2·2H2O]+-Y, and [Mn(EVTCH)2·2H2O]+-Y under mild conditions were studied. These zeolite-Y enslaved Mn(III) complexes are stable and recyclable under current reaction conditions.  相似文献   

18.
The literature concerning the chemical and electrochemical reactions of nitric oxide, nitrous acid and nitrogen dioxide in aqueous solutions is reviewed briefly, with emphasis on electrochemical reductions at platinum electrodes in acidic solutions. The voltammetric behavior of NO and NO2 at a Pt electrode in perchloric acid is virtually identical to that for HNO2 and this is explained on the basis of a common electroactive precursor concluded to be NO+. Three cathodic waves are obtained for acidic solutions of NO, HNO2 and NO2. The first two waves correspond to reduction of NO+ to NO and N2O3 to NO, respectively. The presence of N2O3 results from decomposition of the parent compounds. The presence of Br- or Cl- in acidic solutions of the title compounds promotes the voltammetric reductions at lower H+ concentrations. This probably results from formation of electroactive nitrosyl halides.  相似文献   

19.
The constrained density functional theory (CDFT) was used to investigate the topological effects on intramolecular electron transfer processes that have been reported in previous experimental work [Inorg. Chem., 1997, 36 (22), pp 5037–5049]. The computation mainly focused on three isomers of diferrocenylbenzenes (ortho, para, and meta) and 5-substituted derivatives of m-diferrocencylbenzenes with R = NH2, Cl, CH3, CN, NO2, NeCH3)33+, and N2+. The influence of a third group R′ (R′ = NH2 and N2+) was introduced to the ortho and para isomers. The calculations were compared with the experimental results. The relation between the substituted functional groups and the effectiveness of intramolecular electron transfer was discussed on the basis of CDFT computational results.  相似文献   

20.
Bis(pentamethylcyclopentadienyl)phosphane Cp*2PH reacts with sulfur under basic conditions to give the corresponding dithiophosphinate salts M+ CP*2PS−2 ( 5 M+ = HNEt+3, 6 M+ = Li+), which are formed via the intermediate CP*2P(S)H ( 4 ). Both salts on treatment with cobalt(II) chloride give rise to the transition metal dithiophosphinate 7 . The structures of this new type of diorganodithiophosphinate complexes in the solid state have been investigated. © 1997 John Wiley & Sons, Inc. Heteroatom Chem 8 : 521–525, 1997  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