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.
The energetics of formation of diastereomeric electron-transfer complexes between L-dopa or L-adrenaline and iron(III) complex ions bound to poly(L-glutamate) or poly(D-glutamate) were measured by microcalorimetry at 25°C and pH 7. When the association of substrates by Fe-polypeptide systems was virtualy complete, ΔHll = 1.3 ± 0.1 and ΔHdl = 0.9 ± 0.1 kcal/mol with both substrates. A diastereomeric discrimination energy of 400 ± 100 cal/mol is thus observed, which compares satisfactorily with that directely obtained by differential microcalorimetric measurements. These results are fully consistent with previous findings indicating that thermodynamic effects are of minor importance in the overall stereoselectivity of the electron transfer reactions investigated.  相似文献   

2.
Helix–coil transition of poly(γ-methyl-L -glutamate), poly(γ-ethyl-L -glutamate), and poly(γ-benzyl-L -glutamate) has been studied in mixed solvents by calorimetry, polarimetry, and viscometry. The experimental data have allowed the evaluation of solvation enthalpy Δhb, equilibrium constant K for hydrogen bond formation between the active solvent component and CO and NH groups, and the cooperativity parameter σ. The conformational transition of polypeptides in solution in a mixed solvent containing enough active solvent to maintain the coiled conformation has been produced by dilution with the helix-supporting solvent for the measurements of enthalpy of transition Δhs. The average value for Δhs is 3550 ± 300 J/mol and is practically independent of the nature of the side chain for the dichloroacetic acid-ethylene dichloride solvent pair at 25°C. A noticeable concentration effect exists in the case of poly(γ-benzyl-L -glutamate). The helical conformation is less stable for poly(γ-ethyl-L -glutamate), and this is explained by a steric effect hindering the access of dichloroacetic acid to side chains. Constant K has been calculated using polarimetric data and also from values of Δhs obtained at different temperatures using the Bixon and Lifson theory on the one hand and that of Sayama and coworkers on the other hand. Values of σ for poly(γ-ethyl-L -glutamate) have been calculated according to both theories mentioned, and the results show that the two sets of values are quite similar. The constant σ depends on the nature of the active solvent, on temperature, and on the binary-solvent composition. These conclusions are confirmed by viscometric results. Values of Δhb calculated from constant K are 5230 J/mol when Bixon and Lifson theory is used and 5569 J/mol when the theory at Sayama and coworkers is used. In both cases the value for Δhb is much lower than that of an intramolecular hydrogen bond. Experimental results suggest that the solvation mechanism would proceed in a manner so that mechanisms described in both theories are involved.  相似文献   

3.
A novel flow injection chemiluminescence (CL) method for the determination of rutin was reported. The proposed method was based on the enhanced effect of rutin on the chemiluminescence intensity of luminol and potassium hexacyanoferrate(III) reaction in NaOH medium. The variables of reaction system, such as luminol concentration, potassium hexacyanoferrate(III) concentration and NaOH concentration, were optimized with the aid of response surface methodology. For the responses prediction, a second‐order polynomial model (SOPM) was applied. The optimal conditions for determination of rutin estimated by the model equation were as follows: NaOH concentration of 0.13 mol/L luminol concentration of 0.94 × 10?6 mol/L, and K3Fe(CN)6 concentration of 1.09 × 10?4 mol/L. The theoretical increased ratio of CL intensity (IRI) predicted and actual IRI for 0.05 mg/L rutin under the above conditions were 99.40 and 99.74%, respectively. The SOPM model proved to be powerful for navigating the design space. Under the above optimum conditions, the increased IRI was linearly related to the concentration of rutin in the range from 0.008 to 0.100 mg/L with the regression equation IRI = 1948.20c + 5.24 (r = 0.9994) and in the range from 0.100 to 1.000 mg/L with the regression equation IRI = 1362.50 c + 61.94 (r = 0.9996). The detection limit (3σ) was of 1.95 × 10?3 mg/L. The sampling frequency of this method was 72/h. The method was used directly to determine rutin in tablets. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

4.
The solid-state conformation of copolymers of β-benzyl-L -aspartate [L -Asp(OBzl)] with L -leucine (L -Leu), L -alanine (L -Ala), L -valine (L -Val), γ-benzyl-L -glutamate [L -Glu(OBzl)], or ?-carbobenzoxy-L -lysine (Cbz-L -Lys) has been studied by ir spectroscopy and circular dichroism (CD). The ir spectra in the region of the amide I and II bands and in the region of 700–250 cm?1 have been determined. The results from the ir studies are in good agreement with data obtained by CD experiments. Incorporation of the amino acid residues mentioned above into poly[L -Asp(OBzl)] induces a change from the left-handed into the right-handed α-helix. This conformational change for the poly[L -Asp(OBzl)] copolymers was observed in the following composition ranges: L -Leu, 0–15 mol %; L -Ala, 0–32 mol %; L -Val, 0–8 mol %; L -Glu(OBzl), 3–10 mol %; and Cbz-L -Lys, 0–9 mol %.  相似文献   

5.
《Inorganica chimica acta》1987,133(2):281-287
A series of iron(III)-selective chelating resins have been modeled after the structural features of the naturally occurring siderophore compounds with hydroxamate, catecholate and salicylate iron binding groups. Amberlite IRC-50 was derivatized via an acid chloride intermediate to produce poly(hydroxamic acid) (IRC-50 PHA/N-H and IRC-50 PHA/N- CH3), poly(catecholate) (IRC-50 PEDA 2,3-DHBAD and IRC-50 PEDA 3,4-DHBA) and poly(salicylate) (IRC-50 PEDA 2-HBAD) chelating resins with enhanced iron(III) binding capacity. The poly(hydroxamic acid) IRC-50 PHA/N-CH3 was produced in 57% yield, the highest conversion yet reported for such derivatizations. This is the first report of a catechol or salicylate derivatization of Amberlite IRC-50. The highest overall iron(III) binding capacities yet reported for poly(hydroxamic acid) modifications of commercially available polymer supports were obtained for IRC-50 PHA/N-H (1.75 mmol Fe/g dry resin) and IRC-50 PHA/N-CH3 (1.52 mmol Fe/g dry resin). IRC-50 PHA/N-H was also found to be an effective iron chelator when tested at ambient environmental conditions. Selectivity for iron- (III) was also determined by measuring Fe3+ binding capacity in the presence of Ca2+. UVVis spectroscopy with photoacoustic detection was used to assign the coordination environment of iron(III) in these chelating resins as bis-(FeL2) and tris-(FeL3) chelates. As an alternative synthetic approach, a poly(amidoxime) (PAO) chelating resin was synthesized by polymerization of appropriate monomers. Comparison between the two synthetic approaches showed that PAO bound a greater amount of iron(III) at flow-through column conditions, while IRC-50 PHA/N-H exhibited a greater iron(III) binding capacity at batch equilibration conditions.  相似文献   

6.
《Luminescence》2003,18(3):162-172
The reaction of iron(III) tetrakis‐5,10,15,20‐(N‐methyl‐4‐pyridyl)porphyrin (Fe(III)TMPyP) with hydrogen peroxide (H2O2) and the catalytic activity of the reaction intermediates on the luminescent peroxidation of luminol in aqueous solution were studied by using a double‐mixing stopped‐flow system. The observed luminescence intensities showed biphasic decay depending on the conditions. The initial flashlight decayed within <1 s followed by a sustained emission for more than 30 s. Computer deconvolution of the time‐resolved absorption spectra under the same conditions revealed that the initial flashlight appeared during the formation of the oxo–iron(IV) porphyrin, TMPyPFe(IV) = O, which is responsible for the sustained emission. The absorption spectra 0.0–0.5 s did not reproduce well by a simple combination of the two spectra of Fe(III)TMPyP and TMPyPFe(IV) = O, indicating that transient species was formed at the initial stage. Addition of uric acid (UA) caused a significant delay in the initiation of the luminol emission as well as in the formation of the TMPyPFe(IV) = O. Both of them were completely diminished in the presence of UA equimolar with H2O2, while mannitol had no effect at all. The delay of the light emission as well as the appearance of TMPyPFe(IV) = O was directly proportional to the [UA]0 but other kinetic profiles were not changed significantly. Based on these observations and the kinetic analysis, we confirmed the involvement of the oxo–iron(IV) porphyrin radical cation, (TMPyP)·+Fe(IV) = O, as an obligatory intermediate in the rate‐determining step of the overall reaction, Fe(III)TMPyP + H2O2 → TMPyPFe(IV) = O, with a rate constant of k = 4.3 × 104/mol/L/s. The rate constants for the reaction between the (TMPyP)·+Fe(IV) = O and luminol, and between the TMPyPFe(IV) = O and luminol were estimated to be 3.6 × 106/mol/L/s and 1.31 × 104/mol/L/s, respectively. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

7.
The helix–coil transitions of poly(Nε-methyl, Nε-carbobenzoxy-L -lysine), poly(Nδ-methyl, Nδ-carbobenzoxy-L -ornithine), and poly(Nδ-ethyl, Nδ-carbobenzoxy-L -ornithine) in chloroform–dichloroacetic acid and their corresponding decarbobenzoxylated polypeptides in alkaline solutions were followed by optical rotation measurements. The introduction of a methyl or an ethyl group to the side chains of the carbobenzoxy derivatives of poly(L -lysine) and poly(L -ornithine) appeared to weaken the helical conformation at 25°C. The thermodynamic quantities of the three water-soluble polypeptides were calculated from the data on potentiometric titrations at several temperatures. For uncharged coil-to-helix transition, ΔH = ?370 cal/mol and ΔS = ?1.1 eu/mol for poly(Nε-methyl-L -lysine), and ΔH = ?540 cal/mol and ΔS = ?1.6 eu/mol for poly(Nδ-ethyl-L -ornithine) (all on molar residue basis). The absolute values of ΔH and ΔS dropped in the region of pH-induced transition and eventually both quantities became positive. The initiation factor σ was about 2 × 10?3, which was essentially independent of temperature. For poly(Nδ-methyl-L -ornithine) the coil-to-helix transition was not complete even when the polymer was uncharged at high pH.  相似文献   

8.
Amino acid influx across the brush border membrane of the intact pig ileal epithelium was studied. It was examine whether in addition to system B, systems ASC and bo,+ were involved in transport of bipolar amino acids. The kinetics of interactions between lysine and leucine demonstrates that system bo,+ is present and accessible also to -glutamine. -aspartate (K1/2 0.3 mM) and -glutamate (Ki 0.5 mM) share a high affinity transporter with a maximum rate of 1.3 μmol cm−2 h−1, while only -glutamate with a K1/2 of 14.4 mM uses a low affinity transporter with a maximum rate of 2.7 μmol cm−2 h−1, system ASC, against which serine has a Ki of 1.6 mM. In the presence of 100 mM lysine, -glutamine (A), leucine (B), and methionine (C) fulfilled the criteria of the ABC test for transport by one and the same transporter. However, serine inhibits not only transport of -glutamate but also of glutamine (Ki 0.5 mM), and -glutamate inhibits part of the transport of glutamine. The test does, therefore, only indicate that the three bipolar amino acids have similar affinities for transport by systems B and ASC. Further study of the function of system B must be carried out under full inhibition by lysine and glutamate.  相似文献   

9.
Electron transfer from ortho-dihydroxy substrates, such as L(+)-ascorbic acid, L-adrenaline, and L-dopa, to iron(III) in [Fe(tetpy)(OH)2]+ ions anchored to sodium poly(L-glutamate) (FeTL) or poly(D-glutamate) (FeTD) was found to proceed stereoselectively when structurally ordered and partially shielded active sites prevent easy approach for redox partner. Oxidant-reductant interactions are then mediated by the polypeptide, whose conformational asymmetry ensures an efficient sterically discriminating environment. Evidence is produced that stereoselectivity chiefly arises from transition state effects, while thermodynamic discrimination is of minor importance. Theoretical models of the diastereomeric electron-transfer complexes were constructed by conformational energy calculations based on Coulombic, nonbonded, and hydrogen-bonded energy terms. The molecular parameters of the models enabled "differential" thermodynamic functions of the diastereomeric pairs and stereoselectivity to be evaluated and satisfactorily compared with those experimentally determined. The models give good insight into the observed topochemical phenomena and support the idea that stereoselectivity is coupled with a remote attack mechanism on the central metal ion where the peripheral tetpy ligand of the active sites acts as an electron-transfer agent.  相似文献   

10.
The random copolymerization of the N-carboxyhydrides of γ-benzyl-L -glutamate and L -valine using triethylamine as the initiator in low dielectric media reults in high-molecular-weight copolymers at low convenrson. This behavior makes it possible to apply the monomer reactivity ration theory, which was dervied for addition polymerizations, and from the use of the copolymer composition equation, the respective monomer reactivity ratios, the average and incremental copolymer compositions, and the monomer feed ratio at any conversion can be determined. A comparison of the reactivity ratios for the copolymerization of γ-benzyl-L -glutamate NCA and L -valine NCA in benzene/methylene chloride (rG = 2.1, rV = 0.6) with those obtained using dioxane (rG = 2.7, rV = 0.3) indicates that the interchain compositional heterogeneity is greater for copolymers prepared in the dioxane. For Example, at 100% conversion of the monomeric NCAs, Poly[Glu(OBzl)50Val50] prepared in dioxance has an interchain composition ranging from 74 to 0 mol % γ-benzyl-L -glutamate, whereas in benzene/methylene chloride the interchain composition of γ-benzyl-L -glutamae ranges from 65 to 0 mol %. Once the reactivity ratios are obtained for any pair of α-amino and N-carboxyanhydrides, the use of the aforementioned parameters relating to interchain composition can give insight into the compositional heterogeneity between chains as a function of conversion and provide a basis for the preparation of random α-amino acid copolymers that are homogeneous.  相似文献   

11.
γ-(l)-Menthyl L - and D -glutamates were prepared by a fusion reaction of N-phthalyl-L - and D -glutamic anhydrides with l-menthol, followed by hydrazinolysis. The monomers were polymerized to poly[γ-(l)-menthyl L - and D -glutamates] by the N-carboxyanhydride method. These polymers were soluble in many organic solvents, such as ethyl ether, chloroform, tetrahydrofuran, and n-hexane. From the results obtained by a study of the infrared absorption spectra, the x-ray photographs, the optical rotatory dispersions and the circular dichroisms, poly[γ-(l)-menthyl L -glutamate] was found to be a right-handed α-helix in the solid state and in solution. Similarly, poly[γ-(l)-menthyl D -glutamate] was a left-handed α-helix. The helix-coil transition of these polymers was observed in the vicinity of 40% dichloroacetic acid in a chloroform–dichloroacetic acid mixture.  相似文献   

12.
Late-exponential-phasePenicillium chrysogenum mycelia grown in a complex medium possessed an intracellular iron concentration of 650 μmol/L (2.2±0.6 μmol per g mycelial dry mass). This iron reserve was sufficient to ensure growth and antibiotic production after transferring mycelia into a defined low-iron minimal medium. Although the addition of Fe3+ to the Fe-limited cultures increased significantly the intracellular iron levels the surplus iron did not influence the production of penicillin V. Supplements of purified majorP. chrysogenum siderophores (coprogen and ferrichrome) into the fermentation media did not affect the β-lactam production and intracellular iron level. Neither 150 nor 300 μmol/L extracellular Fe3+ concentrations disturbed the glutathione metabolism of the fungus, and increased the oxidative stress caused by 700 mmol/L H2O2. Nevertheless, when iron was applied in the FeII oxidation state the oxidative cell injuries caused by the peroxide were significantly enhanced.  相似文献   

13.
Burying beetles (Nicrophorus sp.) are necrophagous insects with developed parental care. Genome of Nicrophorus vespilloides has been recently sequenced, which makes them interesting model organism in behavioral ecology. However, we know very little about their physiology, including the functioning of their neuroendocrine system. In this study, one of the physiological activities of proctolin, myosuppressin (Nieve? MS), myoinhibitory peptide (Trica-MIP-5) and the short neuropeptide F (Nicve-sNPF) in N. vespilloides have been investigated. The tested neuropeptides were myoactive on N. vespilloides hindgut. After application of the proctolin increased hindgut contraction frequency was observed (EC50 value was 5.47 x 10-8 mol/L). The other tested neuropeptides led to inhibition of N. vespilloides hindgut contractions (Nicve-MS: IC50 = 5.20 x 10~5 mol/L;Trica-MIP-5: IC50 = 5.95 x 10-6 mol/L;Nicvc-sNPF: IC50 = 4.08 x 10-5 mol/L). Moreover, the tested neuropeptides were immunolocalized in the nervous system of N. vespilloides. Neurons containing sNPF and MIP in brain and ventral nerve cord (VNC) were identified. Proctolin-immunolabeled neurons only in VNC were observed. Moreover, MIP-immunolabeled varicosities and fibers in retrocerebral complex were observed. In addition, our results have been supplemented with alignments of amino acid sequences of these neuropeptides in beetle species. This alignment analysis clearly showed amino acid sequence similarities between neuropeptides. Moreover, this allowed to deduce amino acid sequence of N. vespilloides proctolin (RYLPTa), Nicve-MS (QDVDHVFLRFa) and six isoforms ofNicve-MIP (Nicve-MIP-1一 DWNRNLHSWa;Nicve-MIP-2—AWQNLQGGWa;Nicve-MIP-3—AWQNLQGGWa;Nicve-MlP-4—AWKNLNNAGWa;Nicve-MIP-5—SEWGNFRGSWa;Nicve-MIP-6— DPAWTNLKGIWa;and Nicve-sNPF—SGRSPSLRLRFa).  相似文献   

14.
B. Pispisa  M. Farinella 《Biopolymers》1984,23(8):1465-1479
The oxidation of L -adrenaline (epinephrine) in the presence of [Fe(tetpy)(OH)2]+ ions bound to poly(L -glutamate) or to poly(D-glutamate) has been studied at pH 7 (tetpy = 2,2′:6′,2″:6″,2?-tetrapyridyl). Electron transfer from the substrate to the central metal ion, which is rate-determining, proceeds stereoselectively only when extensive and possibly specific interactions between adrenaline and the peptidic residues of the ordered polymer in the close environment of the active sites occur. This ensures different steric constraints for the two diastereomeric precursor complexes, which are thought to affect the separation and orientation of the redox centers differently, leading to the observed phenomena. Some data on the catalytic oxidation of L -dopa(3,4-dihydroxyphenylalanine) are also presented, showing stereoselective effects similar to those observed with L -adrenaline, despite the diverse distance of the chiral center from the reacting OH groups. A mechanistic interpretation of the results is discussed in the light of a few general considerations concerning the structural features of the catalytic systems. Possible explanations for the finding that stereoselectivity occurs at the expense of the efficiency of catalysis are also considered.  相似文献   

15.
The interaction between synthetic crystalline calcium phosphate (apatite) and acidic macromolecules (sodium polyacrylate, sodium poly(L -glutamate), chondroitin sulfate, phosvitin) was investigated by electron spin resonance spectroscopy of mineral-macromolecule complexes doped with vanadyl ion (VO++) as a paramagnetic probe. Changes in magnetic parameters were interpreted in terms of bonding between mineral and macromolecule. The VO++ probe data indicated that polymer acidic functional groups were bound to mineral surfaces in all cases.  相似文献   

16.
In the first paper heat capacities Cp, of polyglycine, poly(L -alanine), and poly (L -valine) were analyzed using approximate group vibrations and fitting the Cp contributions of the skeletal vibrations to a two-parameter Tarasov function. In this second paper all other poly (amino acid) s are similarly analyzed. Heat capacities were measured by differential scanning calorimetry in the temperature range of 230–390 K for poly(L -leucine), poly(L -serine), poly (sodium-L -aspartate), poly(sodium-L -glutamate), poly(L -asparagine), poly(L -phenylalanine), poly(L -tyrosine), poly(L -methionine), poly (L -tryptophane), poly(L -proline), poly(L -lysine · HBr), poly(L -histidine), poly(L -histidine- HCl), and poly(L -arginine · HCl). Good agreement exists between experiment and calculation. Predictions of heat capacities were made for all not-measured poly (amino acid) s. Enthalpies, entropies, and Gibbs functions for the solid state have been derived. © 1993 John Wiley & Sons, Inc.  相似文献   

17.
Potentiometric, conductometric and 31P NMR titrations have been applied to study interactions between myo-inositol hexakisphosphate (phytic acid), (±)-myo-inositol 1,2,3,5-tetrakisphosphate and (±)-myo-inositol 1,2,3-trisphosphate with iron(III) ions. Potentiometric and conductometric titrations of myo-inositol phosphates show that addition of iron increases acidity and consumption of hydroxide titrant. By increasing the Fe(III)/InsP6 ratio (from 0.5 to 4) 3 mol of protons are released per 2 mol of iron(III). At first, phytates coordinate iron octahedrally between P2 and P1,3. The second coordination site represents P5 and neighbouring P4,6 phosphate groups. Complexation is accompanied with the deprotonation of P1,3 and P4,6 phosphate oxygens. At higher concentration of iron(III) intermolecular P–O–Fe–O–P bonds trigger formation of a polymeric network and precipitation of the amorphous Fe(III)–InsP6 aggregates. 31P NMR titration data complement the above results and display the largest chemical shift changes at pD values between 5 and 10 in agreement with strong interactions between iron and myo-inositol phosphates. The differences in T1 relaxation times of phosphorous atoms have shown that phosphate groups at positions 1, 2 and 3 are complexated with iron(III). The interactions between iron(III) ions and inositol phosphates depend significantly on the metal to ligand ratio and an attempt to coordinate more than two irons per InsP6 molecule results in an unstable heterogeneous system.  相似文献   

18.
Localization of iron-reducing activity in paddy soilby profile studies   总被引:3,自引:0,他引:3  
Profiles of iron speciations (porewaterFe(II) and Fe(III), solid-phase Fe(II) andFe(III)) have been studied to localize both ironreduction and oxidation in flooded paddy soil. Sulfateand nitrate were determined to analyze interactions ofredox reactions involved in the iron cycle with thoseof the sulfur and nitrogen cycle. The development ofthe iron(II) and iron(III) profiles was observed inmicroscale over a time period of 11 weeks. After 11weeks the profiles were stable and showed lowestconcentrations of solid-phase iron(II) on the soilsurface with increasing concentrations to a soil depthof 10 mm ( 100 µmol/cm3). Profilesof iron(III) showed a maximum of iron(III) at a depthof 2 to 4 mm ( 100--200 µmol/cm3).Porewater iron(II) concentrations were three orders ofmagnitude lower than extracted iron(II) and indicatedthat most iron(II) was adsorbed to the solid-phase orimmobilized as siderite and vivianite. Diffusive lossof iron from the soil was indicated by iron recovery(0.3 µmol gdw–1) in the flooding water after12 weeks. The organic content of the soil influencedthe concentrations of solid-phase iron(II) in deepersoil layers (> 6 mm); higher Fe(II) concentrationsin soil with limiting amounts of electron donors mayindicate lower consumption of CO2 by methanogenicbacteria and therefore a higher sideriteprecipitation. Soil planted with rice showed similariron(II) profiles of fresh paddy soil cores. However,maximal iron(III) concentrations ( 350µmol/cm3) were present in planted soil at adepth of 1 to 2.5 mm where oxygen is provided by a matof fine roots. Sulfate and nitrate concentrations inthe porewater were highest on the soil surface (10µM NO3 , 40 µM SO4 2–) anddecreased with depth. Similar profiles were detectedfor malate, acetate, lactate, and propionate, theconcentrations decreased gradually from the surface toa depth of 4 mm. Profiles of oxygen showed highestconcentrations at the surface due to photosyntheticproduction and a depletion of oxygen below 3 mm depth.Methane production rates measured from soil layersincubated separately in closed vessels were zero atthe soil surface and increased with depth. In soildepths below 4 mm where iron(III) concentrationsdecreased higher methane production rates werefound.  相似文献   

19.
The fluorescence of the prulifloxacin (PUFX)–Al(III) system was investigated . Experiments indicated that the fluorescence intensity of prulifloxacin could be greatly enhanced by Al(III) and sensitized by sodium dodecylbenzene sulphonate (SDBS). Accordingly, a sensitive spectrofluorimetric method for the determination of prulifloxacin was established. While excited at 275 nm, the enhanced fluorescence intensity at 412 nm of the system (ΔF) showed a good linear relationship with the concentration of prulifloxacin within the range 4.0 × 10–8–3.0 × 10–6 mol/L. The regression equation was ΔF = 9.83 + 10.8 × 107c (mol/L); the correlation coefficient and detection limit (3σ/k) were 0.99901 and 2.0 × 10–8 mol/L, respectively. The proposed method has been successfully applied to determine prulifloxacin in real pharmaceutical samples. The luminescence mechanism of the system is also discussed in detail. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

20.
The amino-acid-sequence distribution in poly(γ-benzyl-L -glutamate, L -methionine) prepared by polymerization of the respective N-carboxyanhydrides has been investigated. This copolymer was converted first to poly(L -glutamic acid, L -methionine), which was subsequently cleaved by treatment with cyanogen bromide. The resulting material was fractionated into oligomers of (glutamic acid)n-homoserine whose relative molar amounts were determined quantitatively. The results have been compared with those for a random incorporation of the methionine in a γ-benzylglutamate host polymer. Fairly close agreement has been found.  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