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 125 毫秒
1.
The varying degrees of protonation of N-(phosphonomethyl)glycine (PMG, glyphosate) were investigated with infrared (IR) spectroscopy and ab initio frequency calculations. The zwitterionic nature of PMG in solution was confirmed, and intramolecular hydrogen bonding was identified. Successive protonation of the PMG molecule follows the order amine, phosphonate, carboxylate. Intramolecular hydrogen bonding is indicated to exist at all stages of protonation: between both RCO(2-) and RNH(2)(+) and RPO(3)(2-) and RNH(2+) in HL(2)(-) (where L represents the ligand PMG); between RCO(2)(-) and RNH(2)(+) in H(2)L(-); predominantly between RPO(3)(2-) and RNH(2)(+) in H(3)L. There are strong indications that the zwitterion is intact throughout the pH range investigated. Results from IR and extended X-ray absorption fine structure (EXAFS) spectroscopies provide new evidence for structures of N-(phosphonomethyl)glycinecopper(II) complexes. The structures of 1:1 complexes, CuL(-) and CuHL, are essentially the same, differing only in protonation of the phosphonate group. Copper(II) lies at the center of a Jahn-Teller distorted octahedron with all three donor groups (amine, carboxylate, phosphonate) of PMG chelating with copper(II) to form two five-membered chelate rings oriented in the equatorial plane. EXAFS indicates that oxygen (most likely a water molecule) is a fourth ligand, which would thus occupy the fourth corner in the equatorial plane of the elongated octahedron. CuL(2)(4-) most probably forms an isomeric mixture in solution, and there are indications that this mixture is dominated by complexes where two PMG ligands are bound to copper(II) via equatorial and axial positions, with both phosphonate and carboxylate donor groups responsible for chelation at axial positions.  相似文献   

2.
The coadsorption of Cu(II) and glyphosate (N-(phosphonomethyl)glycine, abbreviated to PMG) at the water-goethite interface was studied by means of batch adsorption experiments, attenuated total reflectance Fourier transform infrared (ATR-FTIR) spectroscopy, and extended X-ray absorption fine structure (EXAFS) spectroscopy. The system was investigated over the pH range 3--9 and at total concentrations of 0.9 micromol and 2.2 micromol Cu(II) and PMG per m(2) of goethite. The collective quantitative and spectroscopic results show that Cu(II) and PMG directly interact at the water-goethite interface to form ternary surface complexes. Two predominating complexes have been identified. At pH 4 the IR and CuK-edge EXAFS data indicate a molecular structure where the phosphonate group of PMG bonds monodentately to the surface in an inner sphere mode, while carboxylate and amine groups coordinate to Cu(II) to form a 5-membered chelate ring. Hence, at pH 4, Cu(II) and PMG form a ternary surface complex on goethite with the general structure goethite-PMG-Cu(II). At the highest pH investigated (pH 9), the carboxylate group is still coordinated to Cu(II) but the phosphonate group is present in a relatively free, non-coordinated and/or disordered state. Although the spectroscopic data are not conclusive they indicate the formation of ternary surface complexes with the molecular architecture goethite-Cu(II)-PMG at high pH.  相似文献   

3.
Grignard-reaction of pregnenolone-3β-acetate with pentan-2-one ethylene ketal-5-magnesium bromide and subsequent dehydration yielded 27-nor-25-oxocholesta-5,20 (22)-dien-3β-yl acetate which was hydrogenated to 27-nor-25-oxocholest-5-en-3β-yl acetate and 27-nor-25-oxocholestan-3β-yl acetate.  相似文献   

4.
Polymethacrylylglycinamides (PMG), like polyacrylylglycinamides (PAG), form thermally reversible aqueous gels, but higher molecular weights and/or concentrations are required and the melting points of the gels are lower. The heats of crosslinking for aqueous PMG gels fall in the range of ?5 to ?10 kcal/mole of crosslinks, the same as for aqueous PAG gels, implying that the crosslinks are chemically similar. PMG and PAG are incompatible with each other but both are individually compatible with some types of gelatin. The solubilities of PMG and PAG are similar. Various reagents, however, affect PMG and PAG gels in quite different manners. Aqueous PMG solutions, just outside conditions required for gelation, are rheopectic. Intrinsic viscosities [η] of PMG in 2M NaCNS are about 2.5 times those in water. The Huggins' k′ value for PMG in 2M NaCNS has a value of 0.39–0.40, and both it and [η] are essentially temperature-independent over the range 25–45°C. In water at 25°C for PMG, k′ has an average value of about 1.4. With increasing temperature, for H2O, there is a considerable increase in [η] which is accompanied by a decrease in the value of k′. Osmotic molecular weight measurements on unfractionated PMG in H2O at 40°C yield π/c versus c plots having essentially zero slope, implying a value of close to zero for the second virial coefficient, a value of about 0.5 for the polymer–solvent interaction parameter, and a condition close to a θ condition. An approximate viscosity–M n relationship for polydisperse PMG is [η]2M NaCNS, 25deg;C = 1.7 × 10?8 M n1.5. The low value of K and high value of the exponent do not result from large differences in polydispersity but rather from a stiff, rodlike configuration in solution. This steric hindrance to rotation also manifests itself in the extreme brittleness of PMG films and in a ΔHp for homopolymerization of only ?6 kcal/-mole. The infrared spectra of MG monomer and PMG are recorded as well as the density and refractive index for PMG. PMG has a glass transition at 226°C by DTA and by TGA, thermal decomposition sets in at about 300°C. From copolymerization with acrylic acid, values of 1.66 and +0.06, respectively, were obtained for the resonance factor Q and the electrical factor e for MG monomer.  相似文献   

5.
A novel approach for the simultaneous analysis of glyphosate (PMG), and aminomethylphosphonic (AMPA, GlyP), N-methylaminomethylphosphonic (MAMPA. SarP) and methylphosphonic (MPA) acids is presented. This includes a preliminary 31P NMR analysis of mixtures of PMG, MPA, AMPA and MAMPA, their further derivatization to volatile phosphonates by means of the trifluoroacetic acid-trifluoroacetic anhydride-trimethyl orthoacetate reagent and subsequent MS [chemical ionization (CI) MS, GC-CI-MS, GC-electron impact ionization MS] and/or GC-flame ionization detection (FID) analysis of the products of derivatization. The detection limits of PMG, AMPA, MAMPA and MPA by means of GC-CI-MS and GC-FID were determined. The calibration graphs (GC-FID) for these derivatives were in the range 0.1 to 100 nmol linear and sufficiently reproducible for quantitative determinations. The applicability of the method was demonstrated during the analysis of water samples fortified with PMG, AMPA and MAMPA, characterized by recoveries of >95%.  相似文献   

6.
草甘膦结晶母液经蒸发或纳滤浓缩回收后,产生大量含有高浓度甲醛的废水,常规的废水处理方法难以达到回用或排放要求.以过量溶液浸渍法制备的Pt-Bi-CeO2/AC为催化剂,采用湿式催化氧化法处理2.5%的HCHO溶液,HCHO去除率高达99.9%以上,COD去除率达到96.6%.采用Pt-Bi-CeO2/AC催化剂对含低浓度草甘膦(PMG,50 mg/L)的生产废水直接进行湿式催化氧化处理,催化剂使用23次后,HCHO去除率稳定在85%左右,COD去除率稳定在87%左右,催化剂具有良好的稳定性.湿式催化氧化处理后的废水可直接回用于PMG生产.采用固定床湿式催化氧化装置处理HCHO溶液以及PMG生产废水,处理效果也非常理想,连续使用720 h,催化剂的稳定性能良好.通过XRD、N2吸附-脱附、HRTEM、ICP-OES和XPS等分析手段对催化剂进行了系统表征.  相似文献   

7.
Summary Thermal polymerization of the N-carboxyanhydride of glutamic acid-5-methylester was used to coat microparticulate silica and the surface of capillaries with poly(methylglutamate) (PMG). By increasing the thickness of the PMG layer the hydrophobicity of the stationary phases could be increased. However, total shielding of surface silanols was not achieved. The PMG-coated silicas for HPLC showed unexpected and strange behavior in the chromatography of proteins. Nevertheless, capillaries coated with PMG proved to be excellent for highly efficient CE separations of proteins at medium pH values.  相似文献   

8.
A kinetic study of the polyesterification reaction of γ-phenyl- γ-p-methoxyphenyl-, and γ-p-chlorophenylitaconic acids (1 mole) with ethylene glycol in the presence or in absence of p-toluenesulfonic acid as a catalyst has been carried out in order to show the effect of substituents on the rate and degree of polymerization. The reaction of 1 mole of the acid and an excess of ethylene glycol has also been studied. In all cases the reaction is found to follow the second-order rate equation. The mechanism of polyesterification has been discussed. In catalyzed polyesterification electron-withdrawing groups (CI) decrease the velocity of the reaction. The low values of ρ in both the auto-catalyzed and catalyzed reaction indicate that this polyesterification is slightly sensitive to the polar nature of the substituent.  相似文献   

9.
The viscoelastic behavior of phosphonate derivatives of phosphonylated low-density polyethylene (LDPE) was studied by dynamic mechanical techniques. The polymers investigated contained from 0.2 to 9.1 phosphonate groups per 100 carbon atoms and included the dimethyl phosphonate derivative and two derivatives for which the phosphonate ester group was an oligomer of poly(ethylene oxide) (PEO). The temperature dependences of the storage and loss moduli of the dimethyl phosphonate derivatives were qualitatively similar to those of LDPE. At low phosphonate concentrations, the α, β, and γ dispersion regions characteristic of PE were observed, while at concentrations greater than 0.5 pendent groups per 100 carbons atoms, only the β and α relaxations could be discerned. At low degrees of substitution, the temperature of the β relaxation Tβ decreased from that of PE, but above a degree of substitution of 0.1, Tβ increased. This behavior was attributed to the competing influences of steric effects which tend to decrease Tβ and dipolar interactions between the phosphonate groups which increase Tβ. For the phosphonate containing PEO, a new dispersion region designated as the β′ relaxation was observed as a low-temperature shoulder of the β relaxation. The temperature of the β′ loss was consistent with Tg(U) of the PEO oligomers as determined by differential scanning calorimetry, and it is suggested that the β′-loss process results from the relaxation of PEO domains which constitute a discrete phase within the PE matrix.  相似文献   

10.
A lipoamide (LAm) structure was introduced into a polymeric membrane by chemical modification of poly(γ-methyl-D -glutamate) (PMG). A. redox reaction proceeded across the membrane mediated by pendant LAm groups as solid carriers. It is suggested that the electron transport process in the membrane is derived from the exchange reaction between reduced LAm (thiol) and LAm (disulfide).  相似文献   

11.
A number of highly functionalized dioxocyclams with acetic acid side chains on the secondary amine sites and ethylene glycol side chains on 6 and 13-positions (12a, 12b) or a tetra(ethylene glycol) side chain linking the 6 and 13 positions (15) were synthesized and characterized, as was a bis-dioxocyclam bridged across the 6 and 13 positions by tetra(ethylene glycol) groups (16). These were screened for their ability to complex Gd3+. Only ligands 15 and 16, having tetra(ethylene glycol) groups, formed such complexes.  相似文献   

12.
Radiation-induced terpolymerizations of methyl α,β,β-trifluoroacrylate (MTFA) with tetrafluoroethylene (TFE) and α-olefins, such as ethylene, propylene, and isobutylene, were carried out in bulk at 25°C for the purpose of controlling the content of ester group in the MTFA-α-olefin alternating copolymers. These monomers polymerized to form alternating terpolymers which contained 50 mole % α-olefin in a wide range of monomer composition. The content of MTFA, namely, the ester group in polymer, can be varied without destruction of the alternating structures between fluoroolefins (MTFA, TFE) and α-olefin by changing the MTFA/TFE ratio in the monomer mixture. The relative reactivities of MTFA and TFE in the terpolymerization were discussed according to kinetic treatments by free propagating and complex mechanisms. The relation between the MTFA/TFE ratio in the monomer mixture and that in terpolymer was explained favourbly by the complex mechanism. It was also concluded that the relative reactivity of MTFA is larger than that of TFE in the terpolymerizations.  相似文献   

13.
李波  邓晓军  郭德华  金淑萍 《色谱》2007,25(4):486-490
建立了高效液相色谱-串联质谱测定植物产品(大豆、大米、小麦、蔬菜、水果、茶叶等)、动物肉类产品、水产品、板栗、蜂蜜等产品中草甘膦(PMG)及其主要代谢物氨甲基膦酸(AMPA)残留量的方法。样品经水提取后用二氯甲烷除去其中的脂肪,再经阳离子交换柱(CAX)净化,用 9-芴基甲基氯仿(FMOC-Cl)衍生化,采用多反应监测技术所确定的定性离子对其进行定性,同位素内标法定量。方法的定量检测低限为0.05 mg/kg,线性范围为0.20~10 μg/L,各种基质下PMG和AMPA的平均加标回收率为80.0%~104%,相对标准偏差为6.7%~18.2%。  相似文献   

14.
2′-Deoxy-ψ-isocytidine (VIIβ), a 2′-deoxy analog of antileukemic ψ-isocytidine and also a C-nucleoside analog of deoxycytidine, was synthesized from ψ-uridine by making use of the newly discovered pyrimidine to pyrimidine transformation reaction [J. Chem. Soc., 14, 537 (1977)]. 2′-Deoxy-ψ-uridine (IIβ) and 2′-deoxy-l-methyl-ψ-uridine (V), both C-nucleoside analogs of deoxyuridine and thymidine, were also synthesized. ψ-Uridine was converted into the 2′-chloro analogs (I) which was reduced with tributyltin hydride to give an α,β-mixture of 2′-deoxy-ψ-uridines. The β-isomer (11β was trimethylsilylated and the product (III) treated with methyl iodide to afford the 1-methyl derivative (IV). After hydrolytic removal of the trimethylsilyl groups from IV, the thymidine analog (V) was obtained in good yield. A crude mixture of II was converted in good yield into an α,β-mixture of 1,3-dimethyl-2 -deoxy-ψ-uridines (VI) by treatment with DMF dimethyl acetal in DMF. Treatment of the β-isomer (VIβ) with guanidine, however, gave the α,β-mixture of 2 -deoxy-ψ-isocytidines (VII). The pure β-isomer (VIIβ) was obtained by thick layer chromatography. The pure α-isomer (VIIα) was obtained when VIα was treated with guanidine. 2 -Deoxy-ψ-isocytidine (VIIβ) and 2 -deoxy-l-methyl-ψ-uridine (V) exhibited inhibitory activity against P815 cells (ID5 0 1.2 μg./ml. and 4.9 μg./ml., respectively) and the thymidine analog V was found to be active against Streptococcus faecium var. duran. J. Heterocyclic Chem., 14, 1119 (1977)  相似文献   

15.
Earlier studies have shown that the β-D-digitoxose (2,6-dideoxy-β-D-mannopyranose) directly attached to the cardiac glycoside steroid C3 has the greatest effect on biological activity. This report describes the synthesis of eight digitoxosides (2a-9a), with widely varying cyclic and acyclic C17β-side groups, and the corresponding C3',C4'-acetonides (2b-9b). NMR analysis of conformational strain introduced by the acetonide groups is supported by crystallographic analysis of the sugars' torsion angles.  相似文献   

16.
A series of new 2′–5′ oligonucleotides carrying the 9-(3′-azido-3′deoxy-β-D-xylofuranosyl)adenine moiety as a building block has been synthesized via the phosphotriester method. The use of the 2-(4-nitrophenyl)ethyl (npe) and 2-(4-nitrophenyl)ethoxycarbonyl (npeoc) blocking groups for phosphate, amino, and hydroxy protection guaranteed straightforward syntheses in high yields and easy deblocking lo form the 2′–5′ trimers 21 , 22 , and 25 and the tetramer 23 . Catalytic reduction of the azido groups in [9-(3′-azido-3′-deoxy-β-D-xylofuranosyl)adenine]2′-yl-[2′-(Op-ammonio)→ 5′]-[9-(3′-azido-3′-deoxy-β-D-xylofuranosyl)adenin]-2′-yl-[2′-(Op-ammonio)→ 5′]-9-(3′-azido-3′-deoxy-β-D-xylofuranosyl)adenine ( 21 ) led to the corresponding 9-(3′-amino-3′-deoxy-β-D-xylofuranosyl)-adenine 2′–5′ trimer 26 in which the two internucleotidic linkages are formally neutralized by intramolecular betaine formation.  相似文献   

17.
An efficient bromination protocol for the synthesis of α-bromo-β-keto esters has been developed. In PEG-400 (poly(ethylene glycol-400)), a variety of β-keto esters were treated with NBS (N-bromosuccinimide) at room temperature to selectively afford the corresponding α-monobromination products in excellent yields. It is noteworthy that the reaction was conducted under mild, environmentally benign and catalyst-free conditions.  相似文献   

18.
The fluorescence properties of polymethylene bis-β-naphthoates with various chain lengths (Bn) in poor solvents such as dimethyl sulfoxide-water (DMSO–H2O) and ethylene glycol-water(EG–H2O) were measured under stationary and nonstationary conditions at various temperatures. Hydrophobic interactions bring the two chromophores of B3, B4 and B10 into sandwich arrangement in ground state, thus promote intramolecular excimer formation upon excitation. The perfect overlapping structure of B2 can not be formed in ground state because of the eclipsed conformation of two methylene groups in such a sandwich arrangement. The favorable conformation of B2 is the structure in which the two naphthyl rings are in proximity. The activation energy, enthalpy and entropy changes of excimer formation of B2 were calculated from the fluorescence decay parameters.  相似文献   

19.
The inhibition effect of N,N′-phosphonomethylglycine (PMG) and vinyl phosphonic acid (VPA) on the 3% NaCl acidic solution corrosion of carbon steel iron was studied at different immersion times by potentiodynamic polarization, electrochemical impedance spectroscopy, attenuated total reflectance Fourier transform infrared (ATR-FTIR) spectroscopy, and computational methods. It is found from the polarization studies that PMG and VPA behave as mixed-type inhibitors in NaCl. Values of charge transfer resistance (Rct) and double layer capacitance (Cdl) in the absence and presence of inhibitors are determined. The PMG and VPA inhibitors were capable of inhibiting the corrosion process up to ≈91% and ≈85%, respectively. In the presence of PMG, the synergic effect of chlorine ions was observed. Density functional theory (DFT) was engaged to establish the adsorption site of PMG, VPA, and their deprotonated states. For studied compounds, the resulted values of ELUMO, EHOMO, energy gap (∆E), dipole moment (μ), electronic hardness (η), global softness (σ), electrophilic index (ω), and the electronic potential map are in concordance with the experimental data results regarding their corrosion inhibition behavior and adsorption on the metal surface.  相似文献   

20.
3-Dcazacytosine (4-amino-2-pyridone, 3 ), 3-doazauracil (4-hydroxy-2-pyridone, 5 ), 3-deaza-cytidine (4-amino-1-β-D-ribofuranosyl-2-pyridonc, 9 ), and 3-deazauridine (4-hydroxy-1-β-D-ribo-furanosyl-2-pyridone, 11 ) were prepared in high overall yields from 1-methoxy-1-buten-3-yne ( 1 ). Ethyl 3,5,5-triethoxy-3-pentenoate ( 2 ), obtained from acylatioti of 1 with diethyl carbonate and subsequent in situ conjugate addition of ethoxide, was cyelized with ammonia to provide 3 . Diazotization of 3 and subsequent in situ hydroxydediazotization afforded 5 . Nucleoside 9 was obtained from the stannic chloride-catalyzed condensation of bis-trimethylsilylated 3 and 1-O-acetyl-2,3,5-tri-O-benzoyl-β-D-ribofuranose ( 7 ), followed by ammonolysis of the blocking groups. Diazotization of 9 and subsequent in situ hydroxydediazotization afforded nucleosidc 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