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 kinetics of oxidation of d-erythrose and dL-glyceraldehyde by chromium (VI) and vanadium(V) in perchloric acid medium have been investigated spectrophotometrically. Each reaction was first-order with respect to [oxidant] and [substrate]. The reactions were catalysed by acid, but their dependence on acidity was complex. Sodium perchlorate accelerated the rate of each reaction. The oxidation rates follow the order glyceraldehyde > erythrose. The activation parameters were calculated and mechanisms consistent with the experimental observations are proposed.  相似文献   

2.
《Carbohydrate research》1986,145(2):193-200
The second-order rate constants for the oxidations of d-glucopyranose 6-phosphate and d-ribofuranose 5-phosphate by vanadium(V) in perchloric acid media have been measured spectrophotometrically in the visible region. The order with respect to [H+] is less than unity for each reaction. The enthalpy and entropy of activations of the reactions are higher than those for the corresponding two-electron oxidations by chromium(VI). The oxidation involves free-radical intermediates.  相似文献   

3.
Phytoremediation is an efficient method for the removal of heavy metals from contaminated systems. A productive disposal of metal accumulating plants is a major concern in current scenario. In this work, Cr(VI) accumulating Tradescantia pallida plant parts were investigated for its reuse as a biosorbent for the removal of Cr(VI) ions. The effect of pH, contact time, sorbent dosage, Cr(VI) concentration and temperature was examined to optimize these process parameters. Results showed that Cr(VI) exposed/unexposed T. pallida leaf biomass could remove 94% of chromium with a sorption capacity of 64.672 mg g?1. Whereas the kinetics of Cr(VI) biosorption was well explained by the pseudo second-order kinetic model, the Langmuir model better described the data on Cr(VI) sorption isotherm compared with the Freundlich model. The changes in the free energy (ΔG°), entropy (ΔS°) and enthalpy (ΔH°) were found to be ?5.276 kJ mol?1, 0.391 kJ mol?1 K?1 and 11.346 kJ mol?1, respectively, which indicated the process to be spontaneous, feasible and endothermic in nature. FTIR spectra of T. pallida leaf biomass revealed the active participation of ligands, such as ?NH, amide, hydroxyl and sulphonate groups present in the biomass for Cr(VI) binding, SEM analysis revealed a porous structure of the biosorbent for an easy uptake of Cr(VI).  相似文献   

4.
A partition chromatographic procedure utilizing a cationic exchange resin column in the Li+ form and 90% ethanol as the mobile phase was employed to quantify 3-deoxy-d-manno-octulosonic acid (KDO) and l-glycero-d-manno-heptose in the lipopolysaccharides (LPS) of Re and RdP? rough mutants of Salmonella minnesota. In a standard mixture of monosaccharides, KDO eluted shortly after the void volume and heptose eluted after the neutral hexoses. Mild acid treatment of either the Re or RdP? LPS with 0.16 n methanesulfonic acid in the presence of Dowex 50-X8 resin (H+ form) released more than 80% of the KDO residues within 15 min. The heptose of the RdP? LPS, first detected after 90 min of hydrolysis, increased gradually to a maximum level at 12 h. A secondary gradual increase in KDO became apparent during the heptose release. The weight contents of these two monosaccharides based upon aheir maximum values detected during hydrolysis were 20.3 ± 0.6% KDO, for the Re LPS, and 13.8 ± 0.4% KDO and 12.0 ± 0.4% heptose, for the RdP? LPS. The relationship between the kinetics of release of KDO and heptose and the nature of the linkages involving these two monosaccharides are discussed.  相似文献   

5.
6.
Cyclic dipeptide cyclo(l- or d-Glu-l-His) carrying an anionic site and a nucleophilic site has been synthesized and used as a catalyst for the solvolysis of cationic esters in aqueous alcohols. In the solvolysis of 3-acyloxy-N-trimethylanilinium iodide (S+n, n = 2 and 10) and Cl?H3N+(CH2)11COOPh(NO2), no efficient nucleophilic catalysis was observed. On the other hand, in the solvolysis of Gly-OPh(NO2)·HCl, Val-OPh(NO2)·HCl and Leu-OPh(NO2)·HCl a very efficient general base-type catalysis by cyclo(l-Glu-l-His) was observed. In particular, with the latter two substrates the catalysis by cyclo(l-Glul-His) was more efficient than that by imidazole, although the catalysis was not enantiomer-selective. The diastereomeric cyclic dipeptide cyclo(d-Glu-l-His) was almost inactive under the same conditions. Confomation of cyclo(l- or d-Glu-l-His) in aqueous solution was investigated and the structure/catalysis relationship is discussed.  相似文献   

7.
A mathematical treatment and an original microcalorimetric method are developed to verify an eventual competitive binding between any two substances for the same macromolecule. To apply this method, a competitive binding of L-tryptophan and one benzodiazepin (dipotassium chlorazepate) for human serum albumin is perfectly demonstrated.The association constants and the enthalpy variations are equal to 14 000 ± 2000 M?1 and ?6.6 ± 0.2 kcal/mol for human serum albumin · tryptophan complex and 13 000 ± 1000 M?1 and ?10.0 ± 0.2 kcal/mol for human serum albumin · chlorazepate complex. In all cases the stoichiometry is equal to one.The binding of tryptophan to human serum albumin is partially stereospecific; the association constant and the enthalpy variation for D-tryptophan complex are equal, respectively, to 1000 ± 200 M?1 and ?2.6 ± 0.3 kcal/mol.  相似文献   

8.
The metabolism of D-[1-14C]glucose, D-[6-14C]glucose, D-[1-14C]fructose and D-[6-14C]fructose by leafy spurs of Itea plants results in rapid incorporation of label into allitol and D-allulose. The patterns of labelling found in the allitol and D-allulose are discussed, a direct interconversion from D-glucose and D-fructose being indicated. Allitol has been found to be an active metabolite in Itea plants.  相似文献   

9.
Transport of l-proline into Saccharomyces cerevisiae K is mediated by two systems, one with a KT of 31 μM and Jmax of 40 nmol · s?1 · (g dry wt.)?1, the other with KT > 2.5 mM and Jmax of 150–165 nmol · s?1 · (g dry wt.)?1, The kinetic properties of the high-affinity system were studied in detail. It proved to be highly specific, the only potent competitive inhibitors being (i) l-proline and its analogs l-azetidine-2-carboxylic acid, sarcosine, d-proline and 3,4-dehydro-dl-proline, and (ii) l-alanine. The other amino acids tested behaved as noncompetitive inhibitors. The high-affinity system is active, has a sharp pH optimum at 5.8–5.9 and, in an Arrhenius plot, exhibits two inflection points at 15°C and 20–21°C. It is trans-inhibited by most amino acids (but probably only the natural substrates act in a trans-noncompetitive manner) and its activity depends to a considerable extent on growth conditions. In cells grown in a rich medium with yeast extract maximum activity is attained during the stationary phase, on a poor medium it is maximal during the early exponential phase. Some 50–60% of accumulated l-proline can leave cells in 90 min (and more if washing is done repeatedly), the efflux being insensitive to 0.5 mM 2,4-dinitrophenol and uranyl ions, to pH between 3 and 7.3, as well as to the presence of 10–100 mM unlabeled l-proline in the outside medium. Its rate and extent are increased by 1% d-glucose and by 10 μg nystatin per ml.  相似文献   

10.
An electron diffraction study was carried out on thin single micro-crystals of l-type and dl-type dipalmitoyl lecithins grown in xylene suspensions and fine net patterns were obtained and the mechanism of the thermotropic phase transitions of them was clarified.From the apparent structure of diffraction patterns in low temperature, it is confirmed that the two dimensional lattices have p mm symmetry in l-type and in dl-type lecithins. Lattice parameters from the [001] projection are d100 = 9.9 A? and d010 = 8.8 A? in l-type, and d100 = 17.2 A? and d010 = 8.9 A? in dl-type.With anisotropic variation of dimensions along a and b axes, i.e. contraction for a and expansion for b, induced by temperature rise by electron irradiation during the observation, these diffraction patterns of the lattices of l-type and dl-type were transformed into those characterized by the six diffraction spots having nearly the same spacings. Four of them are observed on slightly outer and two are slightly inner positions as compared with their mean spacings of about (4.1 Å)?1 in l-type and about (4.2 Å)?1 in dl-type. The changes in the patterns observed indicate that at low temperatures the hydrocarbon chains are nearly perpendicular to the layer in dl-type lipid, and tilted with a more complicated packing in l-type ones. The dimension along a in dl-type is twice as large as that in l-type.  相似文献   

11.
A methodological study has been made with a syringe titration unit attached to an LKB batch microcalorimeter. The presicion and accuracy of the instrument assembly have been evaluated by neutralization reactions and by dilution of sucrose solutions. As an example, heat quantities on the order of 10 mJ accompanying the addition of 10 μl titrant solution could be determined with an accuracy of better than 1%. A stepwise titration procedure was used to characterize the binding of indole-3-propionic acid to α-chymotrypsin. The following thermodynamic data were obtained (25°C, acetate buffer, pH 5.80): ΔG0 = ?18.46±0.17 kJ·mol?1, ΔH0 = ?15.26±0.20 kJ·mol?1, ΔS0 = 10.85±1.21 JK?·mol?1.  相似文献   

12.
The anomeric composition and mutarotation rates of fructose 1,6-bisphosphate were determined in the presence of 100 mm KCl at pH 7.0 by 31P NMR. At 23 and 37 °C the solution contains (15 ± 1)% of the α anomer. The anomeric rate constants at 37 °C are (4.2 ± 0.4) s?1 for the β → α anomerization and (14.9 ± 0.5) s?1 for the reverse reaction. A D2O effect between 2.1 and 2.6 was found. From acid base titration curves it appeared that the pK values of the phosphate groups range from 5.8 to 6.0. Mg2+ and Zn2+ bind preferentially to the 1-phosphate in the α-anomeric position. Zn2+ has a higher affinity for this phosphate group than Mg2+ has. At increasing pH the fraction α anomer decreases slightly. At increasing Mg2+/fructose 1,6-bisphosphate ratios the fraction α anomer increases till 19% at a ratio of 20. Proton and probably Mg2+ binding decreases the anomerization rate. The time-averaged preferred orientation of the 1-phosphate along the C1O1 bond of the α conformer is strongly pH dependent, gauche rotamers being predominant at pH 9.4. In the presence of divalent cations the orientation is biased toward trans. A mechanistic model is proposed to explain the Zn2+, Mg2+, and pH-dependent behavior of the gluconeogenic enzyme fructose 1,6-bisphosphatase.  相似文献   

13.
The complexation reactions of O-phospho-DL-serine with Ni(II) or Co(II) were studied at 25°C and ionic strength 0.2 M (KNO3) by temperature-jump. The observed rate constants for formation of the Ni2+ and Co+2 monocomplexes were (1.32 ± 0.09) × 105 and (1.73 ± 0.33) × 107 M?1 sec?1, respectively. Complexation is postulated to involve formation of a monocoordinated steady state intermediate followed by rate-determining chelate ring closure.  相似文献   

14.
The effects of d-glucose addition to a glucose-free luminal perfusate were investigated in the proximal tubule of Necturus kidney, by electrophysiological techniques. The main findings are: (1) In the presence of sodium, d-glucose produces 10.5 mV ± 1.1 (S.E.) depolarization. (2) Phlorizin reduces the magnitude of this response to 2.1 ± 0.1 mV. (3) The glucose-evoked depolarization, ΔVG, does not alter the intracellular K+ activity nor is it affected by peritubular addition of ouabain. (4) Isosmotic reduction of Na+ concentration in luminal perfusate from 95 to 2 mmol/l (choline or Li+ substituting for Na+) does not change the magnitude of ΔVG; complete removal of sodium from the lumen lowers the value of ΔVG (3.2 ± 0.2 mV) but the response is not abolished. This observation suggests that the d-glucose carrier of renal tubules in Necturus is poorly specific with regard to the cotransported cation species.  相似文献   

15.
The kinetics of oxidations of -glucose and -ribose by chromium(VI) and vanadium(V) in perchloric acid medium have been studied. Each reaction was first order with respect to [oxidant] and [substrate]. The reactions were catalysed by acid, but their dependence on acidity was complex. Sodium perchlorate accelerated the rate of reaction. The rate of oxidation of ribose was greater than that of glucose. Mechanisms for these oxidations are suggested. An attempt has also been made to correlate rate constants and activation parameters for the oxidations of different aldoses by these two oxidants.  相似文献   

16.
The title compound, when recrystallised from water, is monoclinic, space group P21, with a = 5.774(4), b = 7.189(5), c = 12.69(1) Å, β = 106.66(5)°, and Z = 2. The crystal structure was determined from three-dimensional X-ray diffraction data taken on an automatic diffractometer with CuKα, and refined by least-squares techniques to R = 0.034 for 977 reflexions. The pyranose ring adopts the 4C1 conformation. The conformation about the exocyclic C-5-C-6 bond is gauche-trans [the torsion angles O-6-C-6-C-5-O-5 and O-6-C-6-C-5-C-4 are 64.2(8) and ?175.6(7)°, respectively], which is significantly different from the gauche-gauche geometry in d-glucose 6-(barium phosphate). The phosphate ester bond, P-O-6, is 1.584(3) Å. All of the oxygen-bonded hydrogen atoms are involved in intermolecular hydrogen-bonds.  相似文献   

17.
Sorbose and 2-deoxy-d-galactose are taken up in Saccharomyces fragilis by an active transport mechanism, as indicated by the energy requirement of the process and the accumulation of free sugar against the concentration gradient. There are no indications for transport-associated phosphorylation as mechanism of energy coupling with these two sugars.The measured sugar-proton cotransport and the influx inhibition by uncouplers suggest a chemiosmotic coupling mechanism. Thus there are at least two different active transport mechanisms operative in Saccharomyces fragilis: transport-associated phosphorylation in the case of 2-deoxy-d-galactose and chemiosmotic coupling in the case of sorbose and 2-deoxy-d-galactose. The difference between the two mechanisms are discussed.Uncouplers do not stimulate downhill sorbose transport in energy-depleted cells and evoke an almost complete inhibition of efflux and of exchange transport.The differences between this sugar-proton cotransport system and similar systems in bacteria and Chlorella are discussed.  相似文献   

18.
Abstract

This research is focussed on kinetic, thermodynamic and thermal inactivation of a novel thermostable recombinant α-amylase (Tp-AmyS) from Thermotoga petrophila. The amylase gene was cloned in pHIS-parallel1 expression vector and overexpressed in Escherichia coli. The steady-state kinetic parameters (Vmax, Km, kcat and kcat/Km) for the hydrolysis of amylose (1.39?mg/min, 0.57?mg, 148.6?s?1, 260.7), amylopectin (2.3?mg/min, 1.09?mg, 247.1?s?1, 226.7), soluble starch (2.67?mg/min, 2.98?mg, 284.2?s?1, 95.4) and raw starch (2.1?mg/min, 3.6?mg, 224.7?s?1, 61.9) were determined. The activation energy (Ea), free energy (ΔG), enthalpy (ΔH) and entropy of activation (ΔS) at 98?°C were 42.9?kJ mol?1, 74?kJ mol?1, 39.9?kJ mol?1 and ?92.3 J mol?1 K?1, respectively, for soluble starch hydrolysis. While ΔG of substrate binding (ΔGE-S) and ΔG of transition state binding (ΔGE-T) were 3.38 and ?14.1?kJ mol?1, respectively. Whereas, EaD, Gibbs free energy (ΔG*), increase in the enthalpy (ΔH*) and activation entropy (ΔS*) for activation of the unfolding of transition state were 108, 107, 105?kJ mol?1 and ?4.1 J mol?1 K?1. The thermodynamics of irreversible thermal inactivation of Tp-AmyS revealed that at high temperature the process involves the aggregation of the protein.  相似文献   

19.
The kinetics of chromium(VI) reduction by Pseudomonas aeruginosa (P. aeruginosa) and Escherichia coli (E. coli) was studied under both pure and mixed cultures. Initially, the study of kinetics was performed in pure culture. It was observed that the growth of the two bacteria was both inhibited in the presence of chromium(VI). The maximum specific growth rate (μ m ) of P. aeruginosa decreased from 2.3942 h?1 (without Cr(VI)) to 1.8551 h?1 (with Cr(VI)). Under the mixed culture, the growth of E. coli was inhibited by P. aeruginosa. The maximum specific growth rate (μ m ) of E. coli decreased form 0.871 h?1 (in pure culture) to 0.153 h?1 (in mixed culture). When the concentration of each bacterium was 4.5 × 108 cells ml?1, the half-velocity reduction rate constant (K C) and the maximum specific reduction rate constant (v max) of chromium(VI) were 80.05 mg chromium(VI) l?1 and 3.674 mg chromium(VI) cells?1 h?1, respectively. The results showed that the simulation appeared in good agreement with the experimental data, supporting the series of mathematical models represented the bacteria growth and chromium(VI) reduction in both pure and mixed cultures usefully.  相似文献   

20.
Thermodynamics of base interaction in (A)n and (A.U)n   总被引:2,自引:0,他引:2  
Using precision scanning microcalorimetry we studied (A)n and (A·U)n melting in highly diluted solutions (0.3 to 5.0 mm) with different Na+ activity. This permitted us to determine directly the thermodynamic functions of stacking interaction in (A)n and base-pairing in (A·U)n. For (A-A) stacking at (A)n melting temperature we obtained ΔH(A)nm = 12.6 kJ mol?1; ΔS(A)nm = 41 J K?1 mol?1. For A·U base-pairing at a standard temperature of 298 K and 0.1 m-Na+ we have: ΔH(A·U) = 34 kJ mol?1; ΔS(A·U) = 102 J K?1 mol?1ΔG(A·U) = ?3.5 kJ mol?1.  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