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.
The rates of formation and dissociation of concanavalin A with some 4-methylumbelliferyl and p-nitrophenyl derivatives of α- and β-D-mannopyranosides and glucopyranosides were measured by fluorescence and spectral stopped-flow methods. All process examined were uniphasic. The second-order formation rate constants varied only from 6.8 · 104 to 12.8 · 104 M?. s?1, whereas the first-order dissociation rate constants ranged from 4.1. to 220 s?1, all at ph 5.0, I = 0.3 M, and 25°C. Dissociation rates thus controlled the value of binding constant. The effect of temperature on these reactions was examined, from which enthalpies and entropies of activation and of reaction could be calculated. The effects of pH at 25°C on the reaction rates of 4-methylumbelliferyl α-D-mannopyranoside and 4-methylumbelliferyl α-D-glucopyranoside with concanavalin A were examined. The value of the binding constant Kap (derived from the kinetics) at any pH could be related to the intrinsic binding constant K by the expression Kap = KaK(Ka + [H+])?1. The values of Ka, the ionization constant of the protein segment responsive to sugar binding, were 3 · 10?4 M and 1 · 10?4 M for 4-methylumbelliferyl α-D-mannopyranoside and 4-methylumbelliferyl α-D-glucopyranoside, respectively. The binding constant of p-nitrophenyl α-D-mannopyranoside is surprisingly much less sensitive to a pH change from 5.0 to 2.7. Ionic strength had little effect on the binding characteristics of 4-methylumbelliferyl α-D-mannopyranoside to concanavalin A at pH 5.2 and 25°C.  相似文献   

2.
The inactivation of Lactobacillus casei thymidylate synthetase by phenylglyoxal occurs by a pseudo-first-order, pH-dependent process which is 100-fold faster at pH 8.4 than at pH 7.4. The second-order rate constant for inactivation at pH 7.4 is 32 m?1 min?1. Although four or more arginyl residues of the 24 arginines per enzyme dimer can be modified, as determined by amino acid analysis or [2-14C]phenylglyoxal incorporation, only one arginine appears to be essential for activity. The association of this arginine with the catalytic process is supported by the finding that 2′-deoxyuridylate not only protects it from modification by phenylglyoxal, but in so doing prevents the enzyme from losing activity. Additional support is derived from the fact that the product of the reaction, 2′-deoxythymidylate, a competitive inhibitor of 2′-deoxyuridylate, also protects the enzyme, but 2′-deoxycytidylate and uridylate do not. Neither the enzyme's second substrate, 5,10-CH2H4folate nor the folylpolyglutamates protect the enzyme from inactivation by phenyglyoxal. These findings contrast with those recently reported by Cipollo and Dunlap (Biochemistry18, 5537, 1979), which indicate that the inactivation is associated with the modification of 4 arginines per mole of enzyme, 2 of which are protected by 2′-deoxyuridylate. The requirement for a single arginine in the catalytic process is consistent with the involvement of one essential cysteine (Noonan et al., Arch. Biochem. Biophys.184, 336, 1977, both amino acids apparently participating in the binding of 1 mol of 2′-deoxyuridylate per enzyme dimer. These findings suggest that the synthetase's two identical subunits function asymmetrically and that 2′-deoxyuridylate binds as a dianion.  相似文献   

3.
The nonenzymatic reduction of nitrosobenzene by NADPH and NADH in aqueous buffer solution at 25°C is described. Both reactants quantitatively convert nitrosobenzene to phenylhydroxylamine. Rate constants for reduction (kr) were determined spectrophotometrically and found to be identical at pH 5.7 and 7.4 and independent of buffer concentration. The values of kNADH (124–149 M?1 sec?1) and kNADPH (131–170 M?1 sec?1) are essentially identical. The reaction is not subject to general catalysis or specific salt effects. The oxidation of phenylhydroxylamine by NAD(P) to nitrosobenzene is only stimulated by a factor of 1.2 over oxidation in its absence (when the ratio of NADP: phenylhydroxylamine was 8:1).  相似文献   

4.
Two esteroproteolytic enzymes (A and D) have been isolated from the mouse submaxillary gland and shown to be pure by ultracentrifugation, immunoelectrophoresis, acrylamide-gel electrophoresis, and amino acid analyses. The enzymes have molecular weights of approximately 30,000 and are structurally and antigenically related. Narrow pH optima between 7.5 and 8.0 are exhibited by both enzymes. The “pK1's” are between 6.0 and 6.5 and the “pK2's” are near 9.0. A marked preference for arginine-containing esters is shown by both enzymes. The maximum specific activity of enzyme A on p-tosylarginine methyl ester (TAME) at pH 8 was 2500–3000 μm min?1 mg?1 and for enzyme D, 400–600 μm min?1 mg?1. With TAME as substrate, the Km for enzyme A was 8 × 10?4m at 25 °C and 6 × 10?4m at 37 °C. For D, Km was 3 × 10?4 at 25 °C and 2 × 10?4m at 37 °C.An apparent activation of enzyme D by tosylarginine (TA), a product of TAME hydrolysis, and all α-amino acids examined was due to removal of an inhibitor by chelation. This effect could be duplicated by 8-hydroxyquinoline and diethyldithiocarbamate but not by EDTA. Enzyme A was not affected by these substances to any remarkable extent. Several divalent ions proved to be potent inhibitors of enzyme D. Both enzymes are inactivated by the active site reagents diisopropyl phosphofluoridate and tosyllysine chloromethylketone but much less rapidly than is trypsin. Nitrophenyl-4-guanidionobenzoate reacts with a burst of nitrophenol liberation but with a rapid continuing hydrolysis. One active site per molecule is indicated. Enzyme D is inactivated by urea, reversibly at 10 m and with maximal permanent losses at 6 m. Autolysis of the unfolded form by the native enzyme when they coexist at intermediate urea concentrations appears to occur.Identity of enzyme D and the epithelial growth factor binding protein is demonstrated.  相似文献   

5.
The aim of this study was to determine acute toxicity in the post larvae of the white shrimp Litopenaeus vannamei after 96 h of exposure to dissolved arsenic under three different temperatures and salinity conditions. Recent reports have shown an increase in the presence of this metalloid in coastal waters, estuaries, and lagoons along the Mexican coast. The white shrimp stands out for its adaptability to temperature and salinity changes and for being the main product for many commercial fisheries; it has the highest volume of oceanic capture and production in Mexican shrimp farms. Lethal concentrations (LC50–96 h) were obtained at nine different combinations (3?×?3 combinations in total) of temperature (20, 25, and 30 °C) and salinity (17, 25, and 33) showing mean LC50–96 h values (±standard error) of 9.13?±?0.76, 9.17?±?0.56, and 6.23?±?0.57 mgAs?L?1(at 20 °C and 17, 25, and 33 salinity); 12.29?±?2.09, 8.70?±?0.82, and 8.03?±?0.59 mgAs?L?1 (at 25 °C and 17, 25, and 33 salinity); and 7.84?±?1.30, 8.49?±?1.40, and 7.54?±?0.51 mgAs?L?1 (at 30 °C and 17, 25, and 33 salinity), respectively. No significant differences were observed for the optimal temperature and isosmotic point of maintenance (25 °C–S 25) for the species, with respect to the other experimental conditions tested, except for at 20 °C–S 33, which was the most toxic. Toxicity under 20 °C–S 33 conditions was also higher than 25 °C–S 17 and 20 °C (S 17 or 25). The least toxic condition was 25 °C–S 17. All this suggests that the toxic effect of arsenic is not affected by temperature changes; it depends on the osmoregulatory pattern developed by the shrimp, either hyperosmotic at low salinity or hiposmotic at high salinity, as observed at least on the extreme salinity conditions here tested (17 and 33). However, further studies testing salinities near the isosmotic point (between 20 and 30 salinities) are needed to clarify these mechanisms.  相似文献   

6.
A thermodynamic study of the binding of flavins (FMN, FAD, 8-carboxylic acid-riboflavin) to the purified apoflavodoxin from Azotobacter vinelandii has been conducted. The binding of FMN was studied at a number of temperatures (10,15, 20, 25, and 30 °C), pH's (6.0, 7.4, and 9.0), and buffer conditions. The binding of FAD was studied at pH 7.4 and 25 °C under a number of buffer conditions. The binding of 8-carboxylic acid-riboflavin to the apoflavodoxin and the binding of FMN to the dimeric form of the apoflavodoxin were investigated at pH 7.4 and 25 °C. Enthalpies of binding for FMN, FAD, and 8-carboxylic- acid-riboflavin were ?28.3, ?16.6, and ?14.0 kcal mol?1, respectively. The enthalpy of binding of FMN to the dimeric form of the apoflavodoxin was ?22.2 kcal mol of binding sites?1. Binding constants of about 108,106, and 106 were obtained for the binding of FMN, FAD, and 8-carboxylic acid-riboflavin, respectively. Using established thermodynamic relationships free energy and entropy changes were calculated. The entropy data indicate that a large degree of ordering of the system occurs upon flavin binding. The pH data suggest that FMN may bind in both the mono-and dianion forms, and that binding doesn't change the pKa of any functional group in the system. It appears that the phosphate group is probably responsible for approximately half the binding enthalpy observed for the binding of FMN. The temperature-dependence data over the temperature range studied is biphasic, centered at 20 °C, indicating that flavin binding occurs to the protein in two thermodynamic states corresponding to the two heat capacities observed. These findings are used to discuss a model for flavin binding.  相似文献   

7.
The oxidation enthalpy of reduced flavin mononucleotide at pH 7.0 in 0.2 m phosphate buffer has been studied by determining the heat associated with the reaction: FMNH2 + 2 Fe(CN)?36 ? FMN + 2 Fe(CN)?46 + 2 H+. (a) (The quinone, semiquinone, and hydroquinone forms of FMN are represented as FMN, FMNH, and FMNH2, respectively.) Calorimetric experiments were performed in a flow microcalorimeter which was modified to prevent sample contamination by oxygen. The enthalpy observed for reaction (a), after correction for dilution and buffer effects, was ?39.2 ± 0.4 kcal (mole FMNH2)?1 at 25 °C. The potential difference, ΔE′, developed by reaction (a) was determined potentiometrically and corresponded to a free energy change, ΔG′, of ?30.3 kcal (mole FMNH2)?1. The resulting entropy change, ΔS′, was thus calculated to be ?29.8 e.u. Reaction (a) was also studied at temperatures of 7 °C and 35.5 °C. ΔCp′ for the reaction was calculated as ?155 ± 18 cal deg?1 (mole FMNH2)?1 at 20 °C. ΔH′ for the reaction (b), FMNH2 ? FMN + H2, (b) was calculated as +14.2 ± 0.7 kcal mole?1 at 25 °C, relative to the enthalpy of the hydrogen electrode being identically equal to zero at all values of pH and temperature. The free energy at pH 7.0 for reaction (b), calculated from the potential was found to be ?9.7 kcal mole?1, which resulted in an entropy for reaction (b) of 80.2 e.u. A thermal titration of reaction (a) was used to calculate the thermodynamic parameters for the formation of semiquinone dimer according to the reaction FMNH2 + FMN ? (·FMNH)2. (c) The free energy, enthalpy, and entropy changes for reaction (c) were estimated to be ?6.1 kcal mole?1, ?7 kcal mole?1, and ?3 e.u., respectively.  相似文献   

8.
Thermoanaerobacter ethanolicus is an extreme thermophilic non-spore forming ethanol-producing anaerobic bacterium. Minimum nutrient requirements and optimum growth conditions have been established. An optimum yeast extract-glucose ratio for ethanol yield has also been determined. Initial medium pH, optimally 7.5–8.0, significantly affected the amount of ethanol formed. Maximum specific growth rate was found to be 0.22 h?1at pH 7.5 and 69°C. Ethanol concentration up to 11 g l?1at pH 7.5 and 69°C was used to characterize ethanol inhibition. The growth kinetics of T. ethanolicus were characterized in terms of environmental parameters. Substrate utilization, ethanol formation and inhibition by both sugar and ethanol were also quantified.  相似文献   

9.
The hydrolyses of p-nitrotrifluoroacetanilide catalyzed by water and imidazole were examined at 70°C. The pH-rate constant profile of the hydrolysis in H2O was examined in the pH range 0.0–11.4. The hydrolysis was independent of pH in the region from pH 1.0 to 4.5, presumably a water-catalyzed reaction. The rate constant and the D2O solvent isotope effect for this reaction were 1.0 × 10?4 sec?1 and 3.7, respectively. Both natural imidazole and imidazolium cation catalyzed hydrolysis. The rate constant of the hydrolysis catalyzed by neutral imidazole was determined to be 5.4 × 10?3M?1 sec?1 and the D2O solvent isotope effect was 1.8.  相似文献   

10.
Bioflocculants of Chlamydomonas reinhardtii were investigated under axenic conditions. C. reinhardtii was found to produce significant amounts of bioflocculants. Flocculating activity by C. reinhardtii began in the linear phase of growth and continued until the end of the stationary phase. The highest flocculating efficiency of the culture broth was 97.06%. The purified C. reinhardtii bioflocculant was composed of 42.1% (w/w) proteins, 48.3% carbohydrates, 8.7% lipids, and 0.01% nucleic acid. The optimum condition for bioflocculant production of C. reinhardtii was as follows: under temperature of 15°C to 25°C, pH 6–10 and illumination of 40–60 μmol photons m?2 s?1. The bioflocculants produced by C. reinhardtii showed maximum activity in pH ranges from 2 to 10. The flocculating activity was significantly enhanced by the addition of CaCl2 as a co-flocculant at an optimal concentration of 4.5 mM.  相似文献   

11.
A low molecular mass cutinase (designated TtcutA) from Thielavia terrestris was purified and biochemically characterized. The thermophilic fungus T. terrestris CAU709 secreted a highly active cutinase (90.4 U ml?1) in fermentation broth containing wheat bran as the carbon source. The cutinase was purified 19-fold with a recovery yield of 4.8 %. The molecular mass of the purified TtcutA was determined as 25.3 and 22.8 kDa using SDS-PAGE and gel filtration, respectively. TtcutA displayed optimal activity at pH 4.0 and 50 °C. It was highly stable up to 65 °C and in the broad pH range 2.5–10.5. Extreme stability in high concentrations (80 %, v/v) of solvents such as methanol, ethanol, acetone, acetonitrile, isopropanol, and dimethyl sulfoxide was observed for the enzyme. The K m values for this enzyme towards p-nitrophenyl (pNP) acetate, pNP butyrate, and pNP caproate were 7.7, 1.0, and 0.52 mM, respectively. TtcutA was able to efficiently degrade various ester polymers, including cutin, polyethylene terephthalate (PET), polycaprolactone (PCL), and poly(butylene succinate) (PBS) at hydrolytic rates of 3 μmol h?1 mg?1 protein, 1.1 mg h?1 mg?1 protein, 203.6 mg h?1 mg?1 protein, and 56.4 mg h?1 mg?1 protein, respectively. Because of these unique biochemical properties, TtcutA of T. terrestris may be useful in various industrial applications in the future.  相似文献   

12.
Biohydrogen production from a simulated fruit wastewater (soluble COD = 3.17 ± 0.10 g L?1) was carried out in a continuous stirred tank reactor (CSTR) of 2 L operational volume without biomass inoculation, heat pre-treatment or pH adjustment, resulting in a low operational pH (3.75 ± 0.09). The hydraulic retention time (HRT) varied from 15 to 5 h. A strong negative correlation (p < 0.01) between the biogas production rate and the HRT was observed. Biogas production rates were higher at 30 °C than at 25 °C (p < 0.01), when the CSTR was operated under the same HRT. The biogas hydrogen content was estimated as high as 55.8 ± 2.3 % and 55.4 ± 2.5 % at 25 and 30 °C, respectively. The main fermentation end products were acetic and butyric acids, followed by ethanol. Significant differences (p < 0.01) during the operation of the CSTR at 25 or 30 °C were identified for butyric acid at almost all HRTs examined. Simulation of the acidogenesis process in the CSTR (based on COD and carbon balances) indicated the possible metabolic compounds produced at 25 and 30 °C reactions and provided an adequate fit of the experimental data.  相似文献   

13.
Formyltetrahydrofolate synthetase (E. C. 6. 3. 4. 3) was found in fresh spinach leaves and purified about 60-fold by treatments of ammonium sulfate, protamine sulfate, dialysis, and DEAE-cellulose column chromatography. Some properties of the enzyme were investigated. Optimum pH was found to be 7.5, and optimum temperature was observed to be at 37°C. In the enzyme reaction, FAH4 and formate were required specifically as the substrates, and Mg++ and ATP were essential components. The Michaelis constants for dl-FAH4, formate, ATP and magnesium chloride were 1.7×10?3 m, 1.7×10?2 m, 4.1×10?4 m and 3.3×10?3 m, respectively. The primary product formed in the reaction catalyzed by the enzyme was suggested as N10-formyl-FAH4 spectrophotometrically. It was observed that the enzyme also catalyzed the reverse reaction. The possible role of the enzyme in plants was discussed.  相似文献   

14.
Mark A. Jensen  Philip J. Elving 《BBA》1984,764(3):310-315
The rate constant, kd, for the dimerization of the free radical (NAD·), produced on the initial one-electron reduction of NAD+, was measured by double potential-step chronoamperometry, fast-scan cyclic voltammetry (cathodic-anodic peak current ratio) and slow-scan cyclic voltammetry (peak potential shift) for a medium in which neither NAD+ nor its reduction products are adsorbed at the solution/electrode interface. All three methods give concordant values of kd (approx. 3·107 M?1·s?1), which are in reasonable accord with the values determined by pulse radiolysis but are considerably greater than values previously determined electrochemically. For the NAD+/NAD· couple, the heterogeneous rate constant (ks,h) exceeds 1 cm·s?1 at 25°C and the formal potential (E0c) vs. sce is ? 1.155 V at 25°C and ? 1.149 V at 1°C at pH 9.1, with an uncertainty of about ±0.005 V.  相似文献   

15.
A recombinant oleate hydratase from Lysinibacillus fusiformis converted ricinoleic acid to a product, whose chemical structure was identified as the novel compound 10,12-dihydroxystearic acid by gas chromatograph/mass spectrometry, Fourier transform infrared, and nuclear magnetic resonance analysis. The reaction conditions for the production of 10,12-dihydroxystearic acid were optimized as follows: pH?6.5, 30 °C, 15 g?l?1 ricinoleic acid, 9 mg?ml?1 of enzyme, and 4 % (v/v) methanol. Under the optimized conditions, the enzyme produced 13.5 g?l?1 10,12-dihydroxystearic acid without detectable byproducts in 3 h, with a conversion of substrate to product of 90 % (w/w) and a productivity of 4.5 g?l?1?h?1. The emulsifying activity of 10,12-dihydroxystearic acid was higher than that of oleic acid, ricinoleic acid, stearic acid, and 10-hydroxystearic acid, indicating that 10,12-dihydroxystearic acid can be used as a biosurfactant.  相似文献   

16.
The performance of the photosynthetic apparatus was examined in the third leaves of Zea mays L. seedlings grown at near-optimal (25 °C) or at suboptimal (15 °C) temperature by measuring chlorophyll (ChI) a fluorescence parameters and oxygen evolution in different temperature and light conditions. In leaf tissue grown at 25 and 15 °C, the quantum yield of PSII electron transport (ψPSII) and the rate of O2 evolution decreased with decreasing temperature (from 25 to 4 °C) at a photon flux density of 125 μmol m?2 s?1. In leaves grown at 25 °C, the decrease of ψPSII correlated with a decrease of photochemical ChI fluorescence quenching (qp), whereas in leaves crown at 15 °C qp was largely insensitive to the temperature decrease. Compared with leaves grown at 25 °C, leaves grown at 15 °C were also able to maintain a higher fraction of oxidized to reduced QA (greater qp) at high photon flux densities (up to 2000 μmol m?2 s?1), particularly when the measurements were performed at high temperature (25 °C). With decreasing temperature and/or increasing light intensity, leaves grown at 15 °C exhibited a substantial quenching of the dark level of fluorescence F0 (q0) whereas this type of quenching was virtually absent in leaves grown at 25 °C. Furthermore, leaves grown at 15 °C were able to recover faster from photo inhibition of photosynthesis after a photoinhibitory treatment (1200 μmol m?2 s?1 at 25, 15 or 6 °C for 8 h) than leaves grown at 25 °C. The results suggest that, in spite of having a low photosynthetic capacity, Z. mays leaves grown at sub optimal temperature possess efficient mechanisms of energy dissipation which enable them to cope better with photoinhibition than leaves grown at near-optimal temperature. It is suggested that the resistance of Z. mays leaves grown at 15 °C to photoinhibition is related to the higher content of carotenoids of the xanthophyll cycle (violaxanthin + antheraxanthin + zeaxanthin) measured in these leaves than in leaves grown at 25 °C.  相似文献   

17.
Carboxysomes are proteinaceous microcompartments that encapsulate carbonic anhydrase (CA) and ribulose 1,5-bisphosphate carboxylase/oxygenase (Rubisco); carboxysomes, therefore, catalyze reversible HCO3 ? dehydration and the subsequent fixation of CO2. The N- and C-terminal domains of the β-carboxysome scaffold protein CcmM participate in a network of protein–protein interactions that are essential for carboxysome biogenesis, organization, and function. The N-terminal domain of CcmM in the thermophile Thermosynechococcus elongatus BP-1 is also a catalytically active, redox regulated γ-CA. To experimentally determine if CcmM from a mesophilic cyanobacterium is active, we cloned, expressed and purified recombinant, full-length CcmM from Nostoc sp. PCC 7120 as well as the N-terminal 209 amino acid γ-CA-like domain. Both recombinant proteins displayed ethoxyzolamide-sensitive CA activity in mass spectrometric assays, as did the carboxysome-enriched TP fraction. NstCcmM209 was characterized as a moderately active and efficient γ-CA with a k cat of 2.0 × 104 s?1 and k cat/K m of 4.1 × 106 M?1 s?1 at 25 °C and pH 8, a pH optimum between 8 and 9.5 and a temperature optimum spanning 25–35 °C. NstCcmM209 also catalyzed the hydrolysis of the CO2 analog carbonyl sulfide. Circular dichroism and intrinsic tryptophan fluorescence analysis demonstrated that NstCcmM209 was progressively and irreversibly denatured above 50 °C. NstCcmM209 activity was inhibited by the reducing agent tris(hydroxymethyl)phosphine, an effect that was fully reversed by a molar excess of diamide, a thiol oxidizing agent, consistent with oxidative activation being a universal regulatory mechanism of CcmM orthologs. Immunogold electron microscopy and Western blot analysis of TP pellets indicated that Rubisco and CcmM co-localize and are concentrated in Nostoc sp. PCC 7120 carboxysomes.  相似文献   

18.
The gene encoding esterase (CE1) from Bacillus pumilus ARA with a calculated molecular weight of 28.4 kDa was cloned, sequenced and efficiently expressed in Escherichia coli. The open reading frame of 747 nucleotides encoded a protein, which was classified as a carboxylesterase with an identity of 87 % to esterase from Bacillus subtilis 168. Recombinant CE1 was purified in a single step to electrophoretic homogeneity by IMAC (Ni2+). The enzyme displayed maximum activity toward p-nitrophenyl (pNP) acetate at 37–40 °C and pH?6.5–7.0. It was stable in the pH range from 6.5 to 8.0, and at temperature from 25 to 40 °C. Among four p-nitrophenyl esters tested, the best substrate was pNP acetate with K m and k cat values of 0.33 mM and 4.07 s?1, respectively. Amounts of 2 mM Ca2+ and Co2+ significantly increased the esterase activity to 190 and 121 %, respectively. These results suggest that CE1 has very attractive applications of increasing feed digestibility in animal nutrition in this moderate temperature range.  相似文献   

19.
We investigated changes in aquatic macrophyte communities over 25 years in Utonai-ko (42°42′N, 141°42′E) in northern Japan and determined the major change-producing factors using canonical correspondence analysis (CCA) of 21 measured hydrochemical variables with potential to influence the occurrence of communities. We then examined the corresponding changes in the 25-year fluctuation trends in the communities and measured variables. The most prominent changes were a decline in the Hippuris community and an increase in the Vallisneria and Myriophyllum communities. CCA revealed that the leading variable was significant wave height (SWH), followed by water depth (WD), total nitrogen (T-N), dissolved oxygen (DO), pH, chlorine ion (Cl?), transparency, mud thickness, and suspended solids (SS). The Hippuris community was positively correlated with T-N and Cl? and negatively with SWH, WD, DO, and pH. All of these variables were likewise correlated with the Vallisneria and Myriophyllum communities but in the reverse direction. SS and transparency exhibited no correlations. During the 25 years, WD and T-N increased, but annual maximum wind velocity and Cl? decreased. Fluctuation of DO was <2 mg ml?1 and pH was consistent. Considering the direction of correlations and 25-year trends, vital factors for the change in aquatic macrophyte communities were WD and Cl?. Because the concentration of C1? was low, the change in aquatic macrophyte communities likely resulted from the increase in WD.  相似文献   

20.
Systematic thermodynamic studies have been conducted for flavin (FMN, FAD) binding to purified riboflavin-binding proteins from hen egg white and egg yolk. These studies were conducted under a variety of temperature (14, 26, and 38 °C), pH (4.5, 5.5, 6.5, 7.4, and 9.0), and buffer conditions, and an extensive thermodynamic profile was constructed. Enthalpies of binding FMN to white riboflavin-binding protein and yolk riboflavin-binding protein were ?19.3 and ?14.4 kcal/mol, respectively, at pH 7.4 and 38 °C. FAD bound to white and yolk riboflavin-binding proteins under the same conditions with ΔH values of ?11.7 and ?6.0, respectively. Binding constants of about 105 and 104 were obtained for FMN and FAD, respectively, and were the same for both proteins under all conditions studied. Using established thermodynamic relationships, we were able to calculate entropy and free energy changes. Entropies indicated a large degree of ordering in the system upon flavin binding with FMN (about ?40 cal/mol/ °C) twice as large as FAD (about ?15 to ?25 cal/mol/ °C), which may indicate a structured solvent interaction with the charged phosphate group, or steric limitations placed on the ribityl side chain in the bound state. Our thermodynamic data support the idea that flavin binding is a mixture of forces, with no one predominant. Analysis of the data suggests that the nucleotide may bind both as the mono- or dianion, that flavin binding occurs with no significant change in the pK of any functional group in the system, except at low pH for FAD binding, and that the temperature variation of the enthalpy change is quite small. These findings are combined with other published data to outline a general scheme of flavin binding with a histidine residue implicated in hydrogen bonding to the adenine portion of FAD, which may be in the unstacked form.  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