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.
Binding of Ca2+ and mg2+ by 2,3-dicarboxyamylose (DCA) and 2,3-dicarboxycellulose (DCC) has been investigated by microcalorimetry, 1H n.m.r. and circular dichroism. Multiple equilibria are apparent between the polyelectrolytes and the cations, with prevalent stoichiometry of one M2+ ion per two monomeric residues (four COO? groups). While complexation of Ca2+ is associated with a conformational transition, no substantial change in the polyelectrolyte conformation is apparent upon binding of Mg2+.  相似文献   

2.
Sphingobium sp. strain SYK‐6 is able to use a phenylcoumaran‐type biaryl, dehydrodiconiferyl alcohol (DCA), as a sole source of carbon and energy. In SYK‐6 cells, the alcohol group of the B‐ring side chain of DCA was first oxidized to the carboxyl group, and then the alcohol group of the A‐ring side chain was oxidized to generate 5‐(2‐carboxyvinyl)‐2‐(4‐hydroxy‐3‐methoxyphenyl)‐7‐methoxy‐2,3‐dihydrobenzofuran‐3‐carboxylate (DCA‐CC). We identified phcF, phcG and phcH, which conferred the ability to convert DCA‐CC into 3‐(4‐hydroxy‐3‐(4‐hydroxy‐3‐methoxystyryl)‐5‐methoxyphenyl)acrylate (DCA‐S) in a host strain. These genes exhibited no significant sequence similarity with known enzyme genes, whereas phcF and phcG, which contain a DUF3237 domain of unknown function, showed 32% amino acid sequence identity with each other. The DCA‐CC conversion activities were markedly decreased by disruption of phcF and phcG, indicating that phcF and phcG play dominant roles in the conversion of DCA‐CC. Purified PhcF and PhcG catalysed the decarboxylation of the A‐ring side chain of DCA‐CC, producing DCA‐S, and showed enantiospecificity towards (+)‐ and (–)‐DCA‐CC respectively. PhcF and PhcG formed homotrimers, and their Km for DCA‐CC were determined to be 84 μM and 103 μM, and Vmax were 307 μmol?min?1?mg?1 and 137 μmol?min?1?mg?1 respectively. In conclusion, PhcF and PhcG are enantiospecific decarboxylases involved in phenylcoumaran catabolism.  相似文献   

3.
John G. Bruno 《Luminescence》1998,13(3):139-145
Electrochemiluminescence (ECL) of 200 ppm 2,3-diaminonaphthalene (2,3-DAN) was studied alone and in conjunction with 100 ppm of 34 different metal and non-metal ions and revealed three relatively intense ECL responses from interactions of 2,3-DAN with Au+, Fe+3 and V+5. ECL responses from Cr+6 or Ru+3 with 2,3-DAN were less intense, but noteworthy, as was the coloured fluorescent product of the non-metal ion Se+4 interaction with 2,3-DAN. Several intense 2,3-DAN–metal ion ECL reactions were studied in greater detail and revealed various titration curves with ionic detection limits in the low ppm range, using a fixed level (200 ppm) of 2,3-DAN. © 1998 John Wiley & Sons, Ltd.  相似文献   

4.
The structures of rat liver and heart plasma membranes were studied with the 5-nitroxide stearic acid spin probe, I(1 2,3). The polarity-corrected order parameters (S) of liver and heart plasma membranes were independent of probe concentration only if experimentally determined low I(1 2,3)/lipid ratios were employed. At higher probe/lipid ratios, the order parameters of both membrane systems decreased with increasing probe concentration, and these effects were attributed to enhanced nitroxide radical interactions. Examination of the temperature dependence of approximate and polarity-corrected order parameters indicated that lipid phase separations occur in liver (between 19° and 28°C) and heart (between 21° and 32°C) plasma membranes. The possibility that a wide variety of membrane-associated functions may be influenced by these thermotropic phase separations is considered. Addition of 3.9 mM CaCl2 to I(1 2,3)-labeled liver plasma membrane decreased the fluidity as indicated by a 5% increase in S at 37°C. Similarly, titrating I(1 2,3)-labeled heart plasma membranes with either CaCl2 or LaCl3 decreased the lipid fluidity at 37°C, although the magnitude of the La3+ effect was larger and occurred at lower concentrations than that induced by Ca2+; addition of 0.2 mM La3+ or 3.2 mM Ca2+ increased S by approximately 7% and 5%, respectively. The above cation effects reflected only alterations in the membrane fluidity and were not due to changes in probe–probe interactions. Ca2+ and La3+ at these concentrations decrease the activities of such plasma membrane enzymes as Na+, K+-ATPase and adenylyl cyclase, and it is suggested that the inhibition of these enzymes may be due in part to cation-mediated decreases in the lipid fluidity.  相似文献   

5.
The structure and reactivity of the complex [Ru(2,3-Medpp)2Cl2](PF6)2 (2,3-Medpp+=2-[2-(1-methylpyridiniumyl)]-3-(2-pyridyl)pyrazine) was investigated by X-ray diffraction (XRD), 1H NMR, redox, and UV-Vis absorption measurements. X-ray analysis shows that crystals obtained from an acetonitrile-toluene solution contain the trans-Cl2, trans-pyrazine isomeric form, while 1H NMR and redox measurements on the main product of the synthetic workup indicate the presence of the trans-Cl2, cis-pyrazine isomer. In the dark at 70 °C, the complex [Ru(2,3-Medpp)2Cl2]2+ reacts slowly in acetonitrile isomerizing to the cis-[Ru(2,3-Medpp)2(CH3CN)Cl]3+ species. Under ambient light in the presence of excess AgNO3 the cis-[Ru(2,3-Medpp)2(CH3CN)2]4+ species is obtained.  相似文献   

6.
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   

7.
Summary Degradation of 3,4-dichloroaniline (34DCA) in aqueous by undefined cultures of free and immobilized cells was examined. Batch cultures of freely suspended cells and continuous degradation in a packed-bed reactor were studied using both synthetically concocted and industrially produced waste-waters. 34DCA was found to be degraded with a concomitant evolution of chloride ions into the bulk medium. The [acked bed reactor with biomass immobilized on celite diatomaceous earth was found to be capable of degrading over 98% of the 34DCA present in a synthetically concocted inlet stream at a concentration of 250 mg l–1. Residence times of less than 4 h were employed, giving an overall volumetric degradation rate for the packed bed of 90 mg l–1 h–1. The industrially produced wastewater contained, in addition to 34DCA, aniline, 4-chloroaniline, 2,3-dichloroaniline (23DCA) and 3,4-dichloronitrobenzene. The biomass enriched on the synthetic 34DCA waste-water was found to be capable of degrading these compounds in addition to 34DCA with the exception of 23DCA. 34DCA degradation efficiencies of over 95% were obtained for the industrial waste-water at reactor residence times of 4.6 h, giving volumetric degradation rates of 24 mg l–1 h–1. Offprint requests to: A. G. Livingston  相似文献   

8.
The meso-2,3-butanediol dehydrogenase (meso-BDH) from S. marcescens H30 is responsible for converting acetoin into 2,3-butanediol during sugar fermentation. Inactivation of the meso-BDH encoded by budC gene does not completely abolish 2,3-butanediol production, which suggests that another similar enzyme involved in 2,3-butanediol formation exists in S. marcescens H30. In the present study, a glycerol dehydrogenase (GDH) encoded by gldA gene from S. marcescens H30 was expressed in Escherichia coli BL21(DE3), purified and characterized for its properties. In vitro conversion indicated that the purified GDH could catalyze the interconversion of (3S)-acetoin/meso-2,3-butanediol and (3R)-acetoin/(2R,3R)-2,3-butanediol. (2S,3S)-2,3-Butanediol was not a substrate for the GDH at all. Kinetic parameters of the GDH enzyme showed lower K m value and higher catalytic efficiency for (3S/3R)-acetoin in comparison to those for (2R,3R)-2,3-butanediol and meso-2,3-butanediol, implying its physiological role in favor of 2,3-butanediol formation. Maximum activity for reduction of (3S/3R)-acetoin and oxidations of meso-2,3-butanediol and glycerol was observed at pH 8.0, while it was pH 7.0 for diacetyl reduction. The enzyme exhibited relative high thermotolerance with optimum temperature of 60 °C in the oxidation–reduction reactions. Over 60 % of maximum activity was retained at 70 °C. Additionally, the GDH activity was significantly enhanced for meso-2,3-BD oxidation in the presence of Fe2+ and for (3S/3R)-acetoin reduction in the presence of Mn2+, while several cations inhibited its activity, particularly Fe2+ and Fe3+ for (3S/3R)-acetoin reduction. The properties provided potential application for single configuration production of acetoin and 2,3-butanediol .  相似文献   

9.
We measured by batch microcalorimetry the standard enthalpy change ΔH° of the binding of Mn2+ to apo-bovine α-lactalbumin; ΔH° = −90 ± 4kJ·mol−1. The binding constants, KMn2+, calculated from the calorimetric and circular dichroism titration curves, are (4.6±1) · 105M−1, respectively. Batch calorimetry confirms the competitive binding of Ca2+, Mn2+ and Na+ to the same site. The relatively small enthalpy change for Mn2+ binding compared to Ca2+ binding favours a model of a rigid and almost ideal Ca2+-complexating site, different from the well-known EF-hand structures. Cation binding to the high-affinity site most probably triggers the movement of an α-helix which is directly connected to the complexating loop.  相似文献   

10.
The interaction of the VO2+ cation with meso-2,3-dimercaptosuccinic acid (DMSA) was investigated by electron absorption spectroscopy in aqueous solution at different pH values. The spectral behavior, complemented with a spectrophotometric titration, shows the generation of a [VO(DMSA)2]2− complex in which the oxocation interacts with two pairs of deprotonated-SH groups of the acid. It was also found that DMSA rapidly reduces VO3 to VO2+, which might be chelated by an excess of the acid. DMSA can also produce the partial reduction of a V2O5 suspension at pH=5.2. The results of this study suggest that DMSA might be a potentially useful detoxification agent for vanadium.  相似文献   

11.
A novel fluorescent probe‐based naphthalene Schiff, 1‐(C2‐glucosyl‐ylimino‐methyl)‐naphthalene‐2‐ol (L) was synthesized by coupling d ‐glucosamine hydrochloride with 2‐hydroxy‐1‐naphthaldehyde. It exhibited excellent selectivity and highly sensitivity for Al3+ in ethanol with a strong fluorescence response, while other common metal ions such as Pb2+, Mg2+, Cu2+, Co2+, Ni2+, Cd2+, Fe2+, Mn2+, Hg2+, Li+, Na+, K+, Fe3+, Cr3+, Zn2+, Ag+, Ba2+ and Ca2+ did not cause the same fluorescence response. The probe selectively bound Al3+ with a binding constant (Ka) of 5.748 × 103 M?1 and a lowest detection limit (LOD) of 4.08 nM. Moreover, the study found that the fluorescence of the L ? Al3+ complex could be quenched after addition of F? in the same medium, while other anions, including Cl?, Br?, I?, NO2?, NO3?, ClO4?, CO32?, HCO3?, SO42?, HSO4?, CH3COO?, PO43?, HPO42?, S2? and S2O32? had nearly no influence on probe behaviour. Binding of the [L ? Al3+] complex to a F? anion was established by different fluorescence titration studies, with a detection limit of 3.2 nM in ethanol. The fluorescent probe was also successfully applied in the imaging detection of Al3+ and F? in living cells.  相似文献   

12.
A 2,3-dihydroxybiphenyl 1,2-dioxygenase from the naphthalenesulfonate-degrading bacterium Sphingomonas sp. strain BN6 oxidized 3-chlorocatechol to a yellow product with a strongly pH-dependent absorption maximum at 378 nm. A titration curve suggested (de)protonation of an ionizable group with a pKa of 4.4. The product was isolated, purified, and converted, by treatment with diazomethane, to a dimethyl derivative and, by incubation with ammonium chloride, to a picolinic acid derivative. Mass spectra and 1H and 13C nuclear magnetic resonance (NMR) data for these two derivatives prove a 3-chloro-2-hydroxymuconic semialdehyde structure for the metabolite, resulting from distal (1,6) cleavage of 3-chlorocatechol. 3-Methylcatechol and 2,3-dihydroxybiphenyl are oxidized by this enzyme, in contrast, via proximal (2,3) cleavage.  相似文献   

13.
Simultaneous uptake of NH+ 4, Na+, Mg2+, Ca2+, NO- 3, SO2- 4, (NO- 2), H2PO- 4 and C1- ions by N-limited winter wheat seedlings(Triticum aestivum L., cv. Regina) in a single depletion experiment was investigated. Individual ion species in uptake solution samples were determined using capillary isotachophoresis. The operating systems used are described in detail. Processing of obtained concentration data allow to construct time curves of individual ions concentration in uptake solution, time curves of accumulative uptake, and curves for uptake rate versus time and uptake rate versus external concentration. From these curves there is possible to reveal a more complex picture of behaviour of individual ions during uptake process as well as to assess interactions among them. Attention was paid to the points of intersection of time curves of NO- 3 and NH+ 4 uptake rates. These points characterized by equal uptake rates may be considered as limits for preference of given ion species under given experimental conditions.  相似文献   

14.
The present study investigates the possible effects of Hg2+, Pb2+, and Cd2+ on [3H]-glutamate binding. To better understand the role of the thiol-disulfide status on the toxicity of such metals toward glutamatergic neurotransmission, we used three thiol chelating agents, 2,3-dimercaptopropanol (BAL), 2,3-dimercaptopropane 1-sulfonate (DMPS), and meso-2,3-dimercaptosuccinic acid (DMSA). Dithiotreitol (DTT) was tested for its ability to prevent metals-induced inhibition on [3H]-glutamate binding. Hg2+, Pb2+, and Cd2+ showed a concentration-dependent inhibition on [3H]-glutamate binding, and mercury was the most effective inhibitor. BAL did not prevent [3H]-glutamate binding inhibition by Hg2+, Cd2+, and Pb2+. However, DMPS and DMSA prevented the inhibition caused by Cd2+ and Pb2+, but not by Hg2+. DTT did not prevent the inhibition on [3H]-glutamate binding caused by 10 M Hg2+. In contrast, it was able to partially prevent [3H]-glutamate binding inhibition caused by 40 M Pb2+ and Cd2+. These results demonstrated that the heavy metals present an inhibitory effect on [3H]-glutamate binding. In addition, BAL was less effective to protect [3H]-glutamate binding inhibition caused by these metals than other chelating agents studied.  相似文献   

15.
Naphthazarin esters (C1–C4) isolated from the roots of Arnebia euchroma are found as skilled dual chemosensors for Ni2+ and Cu2+ among Pb2+, Na2+, K2+, Hg2+, Mg2+, and Ca2+ metal ions. C1–C4 esters exhibited a red shift of 54 nm with Ni2+ and 30 nm with Cu2+ metal ions in absorption. There is a formation of red-shifted bands between 517 and 613 nm in the absorption spectrum of C1–C4 sensors on binding with Ni2+ and Cu2+ ions. The addition of Ni2+ and Cu2+ ions to sensors C1–C4 stimulates a remarkable color change from reddish pink to purple and light blue, respectively. These color changes can be identified with the naked eye. The significant downfield shifts of CO and OH peaks in nuclear magnetic resonance (NMR) spectrum confirm the chelation as binding mechanism. With ultraviolet–visble and NMR studies, it is found that C1–C4 esters possessed notable selectivity and sensitivity toward Ni2+ and Cu2+ over other metal ions.  相似文献   

16.
The results here are the first demonstration of a family of carbohydrate fermentation products opening Ca2+ channels in bacteria. Methylglyoxal, acetoin (acetyl methyl carbinol), diacetyl (2,3 butane dione), and butane 2,3 diol induced Ca2+ transients in Escherichia coli, monitored by aequorin, apparently by opening Ca2+ channels. Methylglyoxal was most potent (K1/2 = 1 mM, 50 mM for butane 2,3 diol). Ca2+ transients depended on external Ca2+ (0.1-10 mM), and were blocked by La3+ (5 mM). The metabolites affected growth, methylglyoxal being most potent, blocking growth completely up to 5 h without killing the cells. But there was no affect on the number of viable cells after 24 h. These results were consistent with carbohydrate products activating a La3+-sensitive Ca2+ channel, rises in cytosolic Ca2+ possibly protecting against certain toxins. They have important implications in bacterial-host cell signalling, and where numbers of different bacteria compete for the same substrates, e.g., the gut in lactose and food intolerance.  相似文献   

17.
13C Nuclear magnetic resonance chemical shifts, 1JC-C scalar coupling constants, spin-lattice relaxation times, and nuclear Overhauser effects were determined for taurine-[1, 2 13C] and a taurine-[1 13C] and taurine-[2 13C] mixture in the presence and absence of calcium. Ionization constants for taurine amino and sulfonic acid groups and chemical shifts of N-methylene and S-methylene carbons of the taurine cation, zwitterion, and anion were obtained from simultaneous least squares analysis of 13C titration curves of both taurine carbons. Comparison of taurine titration shifts to values for related compounds reveals some unusual electronic properties of the taurine molecule. Stability constants of 1:1 calcium complexes with taurine zwitterions and anions, as well as their 13C chemical shifts, were obtained by least squares analysis of titration curves measured in the presence of calcium. The stability constants of calcium-taurine complexes were significantly lower than previous values and led to estimates that only approximately one percent of intracellular calcium of mammalian myocardial cells would exist in a taurine complex. The implications of these results with respect to the effect of taurine on calcium ion flux are discussed.  相似文献   

18.
23Na-NMR investigations of counterion exchange reactions of helical DNA   总被引:2,自引:0,他引:2  
Changes in Δν½, the nmr linewidth of 23Na, have been determined during titrations of helical DNA with polyamines (divalent putrescine and trivalent spermidine) and with inorganic cations (Mg2+ and Co(NH3)). In each case additions of a multivalent cation (Mz+) to a solution containing NaDNA and NaCl cause decreases in Δν½, which is a population-weighted average of contributions from nuclei in bound and free environments. Thus, the binding of Mz+ to DNA displaces sodium ions from regions where the quadrupolar relaxation of 23Na is relatively efficient. At a given extent of titration, the binding of a polyamine produces a smaller decrease in Δν½ than does the binding of an inorganic ion of the same valence. The concentration dependence of Δν½ during the course of a titration can be interpreted most simply as a two-state ion-exchange reaction by assuming that the binding of Mz does not alter RB, the average relaxation rate of sodium nuclei that remain bound. On the basis of this assumption, the initial linear portions of titration curves can be analyzed to determine upper bounds for r°, the number of sodium ions bound per DNA phosphate in the absence of any competing counterion. Analyzing the titration curves for the four multivalent competitors leads to a range of upper-bound estimates for r°: 0.5–0.8. The differences in these estimates could indicate that polyamines displace fewer sodium ions from DNA than do their smaller inorganic counterparts. Alternatively, the range in upper-bound estimates for r° could also reflect specific differences in the effects of the various multivalent cations on RB, if this relaxation rate does change during titration.  相似文献   

19.
The effects of Mg2+, Ca2+, Sr2+, Ba2+, Mn2+, Co2+, Ni2+ and Zn2+ on the kinetics and equilibrium of the association of vacant “tight” ribosomal subunits from Escherichia coli were studied. Increments of Mg2+, Ca2+, Sr2+ and, by and large, Ba2+, to ribosomes dissociated to 30 S and 50 S particles at 1.2 mm-Mg2+ (60 mm-M2+, pH 7.5, 25°C) produce nearly indistinguishable association curves, with midpoints at 1.8 mm total M2+ and complete association to 70 S particles at 4 to 5 mm total M2+ . The association rate constants at 1 mm-Mg2+, 2 mM-M2+ are similar (0.5 × 106 to 0.9 × 106m?1s?1), as are the dissociation rate constants at 1 mm-(Mg2+ + M2+) (0.2 to 0.4 s?1). Mn2+ and Zn2+ increase the degree of association, as well as further aggregation (Zn2+ especially), at lower concentrations than the alkaline earth ions. Co2+ and Ni2+ produce lower degrees of association, by promoting dissociation of the 70 S particle : the association rate constants at 1 mm-Mg2+, 2 mm-M2+ for the transition metal ions are all grouped at 2 × 106 to 3 × 106m?1s?1. Ni2+ also causes a slower inactivation of one or both subunits.The results are compatible with the view that the effects on the rate and equilibrium constants arise from decreases in the electrostatic free energies of the 30 S, 50 S and 70 S particles produced by large-scale, relatively indiscriminate, charge-neutralization “binding” of M2+ , and are difficult if not impossible to reconcile with a specific-sites mode of action of M2+.  相似文献   

20.
Phytase from Nocardia sp. MB 36 was purified (9.65-fold) to homogeneity by acetone precipitation, ion exchange, and molecular sieve chromatography. Native polyacrylamide gel electrophoresis (PAGE) and zymogram analysis showed a single active protein in the purified enzyme preparation. Sodium dodecyl sulfate (SDS)-PAGE analysis showed that phytase was a monomeric protein with a molecular weight of approximately 43 kDa. Phytase exhibited activity and stability over a broad pH range (2–8) and elevated temperatures (50–80°C), and utilized several phosphate compounds as substrates. Phytase was extremely resistant to pepsin and trypsin. Various metal ions viz. Fe2+, Co2+, and Mn2+, and NH4+, ethylenediaminetetraacetic acid or EDTA and phenylmethylsulfonyl fluoride or PMSF had no influence on activity, while Ca2+ and Zn2+ enhanced activity by 15 % and 3.58 %, respectively. SDS caused significant reduction in enzyme activity (41.8 %), while 2,3-butanedione did so moderately (15.9 %). Features of Nocardia sp. MB 36 phytase suggest a potential for animal feed applications.  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