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.
Oxidation products of cytosine, including 5-hydroxycytosine and 5-hydroxyuracil, are highly susceptible to subsequent oxidation. Here, the oxidation products of 5-hydroxy-2'-deoxyuridine have been studied by NMR and MS analyses. The initial products were diastereomers of isodialuric acid nucleoside. These products subsequently decomposed into corresponding dialuric acid derivatives at neutral pH. The position of the carbonyl and hydroxyl groups, at C5 and C6 for isodialuric acid and at C6 and C5 for dialuric acid derivatives, respectively, was determined by 1H- and 13C NMR analyses. In addition, these analyses revealed that the carbonyl groups of both isodaluric and dialuric acid derivatives exist in their fully hydrated form in aqueous solution. Finally, the dialuric acid derivatives were observed to undergo subsequent decomposition into the corresponding 5-hydroxyhydantoin derivatives. Studies of a trinucleotide containing 5-hydroxyuracil suggest that the reactions described herein for the monomer can be extrapolated to DNA.  相似文献   

2.
Ru上有氧条件下氨分解的动力学研究   总被引:3,自引:0,他引:3  
IthasbeenshownthatRuisvalidforthesyn thesisanddecompositionofammonia[1,2 ] .FurtherstudyofammoniaadsorptionanditsdecompositionproductsdesorptiononRuwillbeimportant .Previ ousstudiesofammoniaadsorptiononRumainlyfo cusedontheammoniasynthesisandhydrogenpro ductionintheabsenceofoxygen[3] ,onlyafewinves tigationsonammoniadecompositioninthepresenceofoxygenhavebeenreported[4 ,5] ,andtheeffectofad sorbedoxygenontheratesofammoniadecompositionandproductformationonRuarestillnotwellunder stood .Inthispa…  相似文献   

3.
Activity experiments using GC analysis of reactor effluent have been combined with in situ IR spectroscopy to elucidate the reaction steps in the destructive adsorption of CHCl3, CH2Cl2, and CH3Cl over LaOCl. The IR results show that during reaction, LaOCl is covered with carbonate, formate, and methoxy groups. The relative amount of each of these surface intermediates depends on the Cl/H ratio of the reactant. The decomposition of the surface species leads to formation of the reaction products, and is influenced by the temperature and the relative amount of Cl present on the surface. The GC results show that the activity for the destructive adsorption of H-containing chlorinated C1 compounds decreases with increasing hydrogen content of the reactant. The acquired insight into the mechanism of destructive adsorption is crucial to the design of new catalyst materials for the efficient conversion of chlorinated hydrocarbons into nonhazardous products or reusable chemicals.  相似文献   

4.
Electronic structure calculations have been performed to investigate the initial steps in the gas-phase decomposition of urea and urea nitrate. The most favorable decomposition pathway for an isolated urea molecule leads to HNCO and NH3. Gaseous urea nitrate formed by the association of urea and HNO3 has two isomeric forms, both of which are acid-base complexes stabilized by the hydrogen-bonding interactions involving the acidic proton of HNO3 and either the O or N atoms of urea, with binding energies (D0(o), calculated at the G2M level with BSSE correction) of 13.7 and 8.3 kcal/mol, respectively, and with estimated standard enthalpies of formation (delta(f)H298(o) of -102.3 and -97.1 kcal/mol, respectively. Both isomers can undergo relatively facile double proton transfer within cyclic hydrogen-bonded structures. In both cases, HNO3 plays a catalytic role for the (1,3) H-shifts in urea by acting as a donor of the first and an acceptor of the second protons transferred in a relay fashion. The double proton transfer in the carbonyl/hydrogen bond complex mediates the keto-enol tautomerization of urea, and in the other complex the result is the breakdown of the urea part to the HNCO and NH3 fragments. The enolic form of urea is not expected to accumulate in significant quantities due to its very fast conversion back to H2NC(O)NH2 which is barrierless in the presence of HNO3. The HNO3-catalyzed breakdown of urea to HNCO and NH3 is predicted to be the most favorable decomposition pathway for gaseous urea nitrate. Thus, HNCO + NH3 + HNO3 and their association products (e.g., ammonium nitrate and isocyanate) are expected to be the major initial products of the urea nitrate decomposition. This prediction is consistent with the experimental T-jump/FTIR data [Hiyoshi et al. 12th Int. Detonation Symp., Aug 11-16, San Diego, CA, 2002].  相似文献   

5.
The products of the thermal decomposition of a series of copolymers of structure with n = 0, 1, 3, 4 and 5, have been investigated. A high proportion of the products consists of low molecular weight analogues of the parent polymers, with styrene the major constituent of the volatile liquids. The copolymers are more stable than head-to-tail polystyrene. Some general conclusions are drawn on the overall mechanism of breakdown.  相似文献   

6.
Density functional calculations have been carried out to determine geometries, adsorption energies and vibrational frequencies of NO, N(2)O, CO, O(2), and H(2)O, on a model for Ag(I) and Au(I) ion-exchanged ZSM-5 catalysts. Using statistical mechanics, the DeltaH and DeltaG values were calculated in order to evaluate the stability of the adsorbates on Ag(I) and Au(I) sites. The calculated vibrational frequencies are in reasonable agreement with the reported experimental values. The analysis of the results shows that at 475 degrees C the adsorption of two NO molecules and the direct N(2)O decomposition on AgZSM-5 are thermodynamically unfavorable. The adsorption of one NO molecule presents a small positive DeltaG value. On the contrary, in the case of AuZSM-5, the adsorption of one NO molecule and the direct N(2)O decomposition to produce N(2) are thermodynamically favorable. For both models, the N(2)O decomposition by AgO and AuO species is thermodynamically very favorable. The analysis of the interaction with H(2)O shows that water displaces the adsorbed NO on AgZSM-5 but not on AuZSM-5 which indicates that the AuZSM-5 catalyst is less sensitive to deactivation by H(2)O than the AgZSM-5 catalyst.  相似文献   

7.
Dried samples of polyacrylamide in an He atmosphere have been subjected to thermogravimetric analysis in the 30–600°C range, and the evolved gases were monitored by FTIR. Water, ammonia, and small quantities of carbon dioxide are released in the first stages of decomposition (220–340°C), where the polymer chains remain intact and the reaction occurs on the pendant amide groups. In the second stage of decomposition (340–440°C), the majority of the weight loss occurs, and main chain breakdown occurs, releasing carbon dioxide, water, nitrile compounds, and imides. Trapping of the gases in this stage and analysis by GC–FTIR and GC–MS reveals the presence of more than 20 decomposition products, and confirms that a large proportion of these can be assigned to glutarimide and its substituted analogs. Imidization and dehydration reactions on the amide groups, as well as free radical breakdown of the main chains, with inter- and intramolecular hydrogen transfer, can account for many of the products of the decomposition. © 1993 John Wiley & Sons, Inc.  相似文献   

8.
Several studies in the literature indicate that histones (lysine rich proteins found associated with DNA in eukaryotic chromatin), as well as poly-L-lysine, can be photocross-linked by ultraviolet (UV) light to DNA in which 5-bromo-2'-deoxyuridine has been substituted for thymidine. To gain some insight into the possible nature of this cross-linking, we have studied the photoreactions occurring in deoxygenated aqueous solutions containing 5-bromouracil (I) (BrUra) or 5-bromo-2'-deoxyuridine (III) (BrdUrd) and ethylamine, a lysine side chain analog. In the case of I this reaction produced the ring opened compound N-(N'-ethylcarbamoyl)-3-amino-2-bromoacrylamide (Ia). A small amount of N-(N'-ethylcarbamoyl)-3-ethylamino-2-bromoacrylamide (Ic) was also isolated. It was found that purified Ia, standing in the presence of ethylamine, was gradually converted to Ic in a dark reaction. The beta and alpha anomers of N-(N'-ethylcarbamoyl)-3-(2'deoxyribofuranos-1'-yl) amino-2-bromoacrylamide (IIIa and IIIb respectively) were isolated as products in the photoreaction of III with ethylamine; the alpha anomer was produced in a dark reaction from the beta anomer. The identity of these anomers was established by comparison of their proton NMR spectra with those of the four corresponding alpha and beta furanosyl and pyranosyl isomeric nucleosides of thymine, which are presented in the Appendix. A study was also made of the reaction of I with methylamine; a ring opened product analogous to Ia, viz. N-(N'-methylcarbamoyl)-3-amino-2-bromoacrylamide (IIa) was formed. A similar study with 5-bromo-1-methyluracil produced N-(N'-methylcarbamoyl)-3-methylamino-2-bromoacrylamide (IIc) as a product. Likewise, the reaction of 5-chlorouracil with ethylamine was studied and N-(N'-ethylcarbamoyl)-3-amino-2-chloroacrylamide (Ie), which is analogous in structure to Ia, was found to be produced. Structural identifications were made through use of UV spectroscopy, high resolution 1H-NMR spectroscopy, mass spectrometry and, in the case of Ia and IIa, 13C-NMR spectroscopy. In the BrUra and BrdUrd reaction systems, described above, dehalogenation reactions accounted for a major portion of the products. The yields of ring opened products, determined at pH 10, ranged from a high of 10.3% in the BrUra-ethylamine system to a low of 1.7% in the MeBrUra-methylamine system.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

9.
The adsorption and thermal decomposition of alkanethiols (R-SH, where R = CH3, C2H5, and C4H9) on Pt(111) were studied with temperature-programmed desorption (TPD) and X-ray photoelectron spectroscopy (XPS) with synchrotron radiation. Dissociation of sulfhydryl hydrogen (RS-H) of alkanethiol results in the formation of alkanethiolate; the extent of dissociation at an adsorption temperature of 110 K depends on the length of the alkyl chain. At small exposure, all chemisorbed CH3SH, C2H5SH, and C4H9SH decompose to desorb hydrogen below 370 K and yield carbon and sulfur on the surface. Desorption of products containing carbon is observed only at large exposure. In thermal decomposition, alkanethiolate is proposed to undergo a stepwise dehydrogenation: R'-CH2S --> R'-CHS --> R'-CS, R' = H, CH3, and C3H7. Further decomposition of the R'-CS intermediate results in desorption of H2 at 400-500 K and leaves carbon and sulfur on the surface. On the basis of TPD and XPS data, we conclude that the density of adsorption of alkanethiol decreases with increasing length of the alkyl chain. C4H9SH is proposed to adsorb mainly with a configuration in which its alkyl group interacts with the surface; this interaction diminishes the density of adsorption of alkanethiols but facilitates dehydrogenation of the alkyl group.  相似文献   

10.
Infrared and molecular beam experiments are used to compare and contrast the adsorption and reaction of NO and trans-1,2-dichloroethene on Cu(110) and on Cu nanoclusters deposited on a 5 A thick Al(2)O(3) film. The overall reaction of NO, leading to decomposition, is almost identical in the two systems, with both types of Cu surfaces promoting the formation of NO dimers, which are precursors to the dissociation products N(2)O, N(2) and O. Although the overall reaction is independent of surface structure, the IR spectra clearly show differences in the adsorption sites occupied on the single crystal and the clusters, a disparity that is also shown by CO adsorption experiments. In contrast, the reaction pathway of dichloroethene does show differences on the two types of Cu surfaces. On both surfaces, the initial reaction step is insensitive to structure and efficient dechlorination leads to the production of adsorbed acetylene. However, the fate of this intermediate depends critically on the underlying surface. On Cu(110), the acetylene trimerises readily into benzene at 350 K. However, this reaction shows a significant size dependent behaviour on the supported nanocluster systems, with the probability for trimerisation diminishing with decreasing cluster size.  相似文献   

11.
Hydrolytic reactions of the structural analogue of guanylyl-(3',3')-uridine, guanylyl-(3',3')-(2'-amino-2'-deoxyuridine), having one of the 2'-hydroxyl groups replaced with an amino function, have been followed by RP HPLC in the pH range 0-13 at 90 degrees C. The results are compared to those obtained earlier with guanylyl-(3',3')-uridine, guanylyl-(3',3')-(2',5'-di-O-methyluridine), and uridylyl-(3',5')-uridine. Under basic conditions (pH > 8), the hydroxide ion-catalyzed cleavage of the P-O3' bond (first-order in [OH(-)]) yields a mixture of 2'-amino-2'-deoxyuridine and guanosine 2',3'-cyclic phosphate which is hydrolyzed to guanosine 2'- and 3'-phosphates. Under these conditions, guanylyl-(3',3')-(2'-amino-2'-deoxyuridine) is 10 times less reactive than guanylyl-(3',3')-uridine. Under acidic and neutral conditions (pH 3-8), where the pH-rate profile for the cleavage consists of two pH-independent regions (from pH 3 to pH 4 and from 6 to 8), guanylyl-(3',3')-(2'-amino-2'-deoxyuridine) is considerably reactive. For example, in the latter pH range, guanylyl-(3',3')-(2'-amino-2'-deoxyuridine) is more than 2 orders of magnitude more labile than guanylyl-(3',3')-(2',5'-di-O-methyluridine), while in the former pH range the reactivity difference is 1 order of magnitude. Under very acidic conditions (pH < 3), the isomerization giving guanylyl-(2',3')-(2'-amino-2'-deoxyuridine) and depurination yielding guanine (both first-order in [H(+)]) compete with the cleavage. The Zn(2+)-promoted cleavage ([Zn(2+)] = 5 mmol L(-)(1)) is 15 times faster than the uncatalyzed reaction at pH 5.6. The mechanisms of the reactions of guanylyl-(3',3')-(2'-amino-2'-deoxyuridine) are discussed, particularly focusing on the possible stabilization of phosphorane intermediate and/or transition state via an intramolecular hydrogen bonding by the 2'-amino group.  相似文献   

12.
Thermal decomposition of strontium titanyl oxalate tetrahydrate and calcium titanyl oxalate hexahydrate have been studied employing TG, DTA, gas and chemical analysis. The decompositions proceed through three major steps: dehydration, decomposition of the oxalate to a carbonate and the decomposition of the carbonate to yield the final products, the metatitanates. The intermediates of the oxalate decomposition are found to be Sr2Ti2O4+x(CO3)2-x(CO2)x and Ca2Ti2O4(CO3)2, respectively. The entrapment of carbon dioxide in the former and the presence of non-equivalent carbonate groups in the latter are substantiated by their i.r. spectra. The penultimate solid residues are poorly crystalline Sr2Ti2O5CO3 and amorphous Ca2Ti2O5CO3. Decompositions of these carbonates are accompanied by growth in particle size of the products, SrTiO3 and CaTiO3, respectively.  相似文献   

13.
The infrared spectra of the oxodiperoxo and triperoxo complexes, (NH4)[VO(O2)2(phen)].2H2O and (NH4)[V(O2)3(phen)].2H2O have been recorded and the observed bands are assigned on the basis of Cs symmetry. Thermogravimetric (TG) and differential thermal analysis (DTA) measurements on these two complexes were also carried out. A detailed mechanism for the mode of thermal decomposition of the two complexes has been given and supported by infrared spectral measurements on the thermal decomposition products. The data obtained agree quite well with the expected structure and indicate that the final thermal decomposition product of these two complexes is V2O5.  相似文献   

14.
Reactions of 5-p-tolylethynyl-2'-deoxyuridine and 3',5'-di-O-acetyl-5-p-tolylethynyl-2'-deoxyuridine with Co2(CO)8 in THF gave 5-p-tolC2[CO2(CO)6]-2'-deoxyuridine and 3',5'-di-O-acetyl-5-p-tolC2[CO2(CO)6]-2'-deoxyuridine (92 and 66%).  相似文献   

15.
The methods of temperature-programmed reaction/desorption (TPR/TPD) are used to study azomethane (CH3N=NCH3) decomposition and the reactions of the products of its pyrolysis (CH 3 * radicals and N2) on the polycrystalline molybdenum surface. A TPR spectrum of adsorbed azomethane decomposition shows mainly N2, H2, and unreacted azomethane. Upon preliminary adsorption of azomethane pyrolysis products on a catalyst sample, a TPR spectrum shows N2, H2, and CH4 in comparable amounts. The difference in the composition of desorption products found for these two types of experiments shows that, in the decomposition of adsorbed azomethane, surface methyl moieties are not formed. The rate constants were calculated for the dissociation of adsorbed CH3, CH2, and CH, recombination of hydrogen atoms with each other and with CH3 and CH2, and the recombinative desorption of nitrogen atoms. Deceased.  相似文献   

16.
High-resolution thermogravimetry (HR TG) was used to study the adsorption of mercury(II) ions by modified MCM-41 material and regeneration of the loaded adsorbent with mercury ions by using different eluents. The weight change curves were measured for MCM-41 samples modified with 1-benzoyl-3-propylthiourea ligand loaded with mercury ions. The differential thermogravimetric (DTG) curves were analyzed to investigate the adsorption of mercury ions by the aforementioned multifunctional ligand and to monitor the decomposition of the metal-ligand complexes. A series of experiments performed for different Hg(2+):ligand ratios allowed us to correlate the adsorption data for mercury ions measured by means of UV spectrophotometry with those obtained by HR TG analysis. The DTG results provided additional information about mercury-ligand interactions as well as the thermal stability of mercury-ligand complexes. This study shows that HR TG is a very attractive technique for studying the adsorption of mercury ions on modified nanoporous silicas and monitoring their regeneration. Since the samples used are small, this method seems to be promising for studying adsorption systems of environmental significance.  相似文献   

17.
用TPD和IR方法研究了CH_3NO_2在典型固体酸SiO_2-Al_2O_3和固体碱MgO催化剂上的吸附分解。结果表明,在SiO_2-Al_2O_3表面CH_3NO_2吸附转化为表面甲酰胺物种,后者在高温下分解为CO_2和NH_3。在MgO表面CH_3NO_2吸附形成多种表面化学物种,它们在升温过程中脱附,并通过表面亚硝基甲烷物种分解为NO、C_2H_4、C_2H_6和N_2O.讨论了CH_3NO_2分解过程中表面酸、碱中心的作用。  相似文献   

18.
Ellis DA  Martin JW  Muir DC  Mabury SA 《The Analyst》2003,128(6):756-764
Fluoropolymers are widely used in industry and consumer products. At the thermal limit of their stability (e.g. > 260 degrees C for PTFE) numerous studies have reported a variety of thermolysis products produced upon polymer breakdown. In the current investigations our objective was to expand the knowledge of these products by advancing the techniques used to obtain their identity. The use of 19F NMR to compliment derivatization with GC-MS has been shown to facilitate the identification of novel fluorinated species, in particular fluorinated acids, that had, until recently, gone previously unreported for the thermal decomposition of fluorinated polymers using traditional techniques. The polymers chosen for the decomposition studies were poly(tetrafluoroethylene), poly(chlorotrifluoroethylene), poly(ethylene-chlorotrifluoroethylene) and poly(tetrafluoroethylene-co-tetrafluoroethylene perfluoropropyl ether) which cover the three major classes of industrially produced fluoro-polymer, co-polymer and elastomer. The use of 1D 19F and 2D 19F-19F correlation spectroscopy (COSY) NMR allowed for the observation of polyfluorinated acids and their atmospheric precursors. This in turn allowed the modification of GC-MS procedures to verify these NMR findings. NMR results also showed a plethora of unidentified and previously unreported materials, thermolysis products that await characterization.  相似文献   

19.
The thermal decompositions of methyl azidoformate (N3COOMe), ethyl azidoformate (N3COOEt) and 2-azido-N,N-dimethylacetamide (N3CH2CONMe2) have been studied by matrix isolation infrared spectroscopy and real-time ultraviolet photoelectron spectroscopy. N2 appears as an initial pyrolysis product in all systems, and the principal interest lies in the fate of the accompanying organic fragment. For methyl azidoformate, four accompanying products were observed: HNCO, H2CO, CH2NH and CO2, and these are believed to arise as a result of two competing decomposition routes of a four-membered cyclic intermediate. Ethyl azidoformate pyrolysis yields four corresponding products: HNCO, MeCHO, MeCHNH and CO2, together with the five-membered-ring compound 2-oxazolidone. In contrast, the initial pyrolysis of 2-azido-N,N-dimethyl acetamide, yields the novel imine intermediate Me2NCOCH=NH, which subsequently decomposes into dimethyl formamide (HCONMe2), CO, Me2NH and HCN. This intermediate was detected by matrix isolation IR spectroscopy, and its identity confirmed both by a molecular orbital calculation of its IR spectrum, and by the temperature dependence and distribution of products in the PES and IR studies. Mechanisms are proposed for the formation and decomposition of all the products observed in these three systems, based on the experimental evidence and the results of supporting molecular orbital calculations.  相似文献   

20.
The adsorption and photodecomposition of seven kinds of amino acids on a TiO2 surface were investigated by zeta potential measurements and 1H NMR spectroscopy in TiO2 aqueous suspension systems. The decomposition rates increased in the order of Phe < Ala < Asp < Trp < Asn < His < Ser. For Phe, Trp, Asn, His, and Ser, the isoelectric point (IEP) of TiO2 shifted to a lower pH with increasing decomposition rates upon adsorption on TiO2, suggesting that the effective adsorption and photocatalytic sites for these amino acids should be the basic terminal OH on the solid surface. Since the amino acids that decomposed faster than the others contain -OH (Ser), -NH (Trp, His), or -NH2 (Asn) in their side chain, they are considered to interact with the basic terminal OH groups more preferably by the side chain and are vulnerable to photocatalytic oxidation. On the other hand, Ala interacts with the acidic bridged OH on TiO2 to cause an IEP shift to a higher pH. The correlation of the surface hydroxyl groups with the photocatalysis of amino acids was verified by the use of calcined TiO2 without surface hydroxyl groups.  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