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.
An optimization study has been performed on 2,3-butanediol production by Bacillus licheniformis NCIMB 8059 from different carbon sources (glucose, sucrose and cornstarch hydrolyzate), alternately varying temperature (34相似文献   

2.
The production of xylitol from concentrated synthetic xylose solutions (S(o) = 130-135 g/L) by Debaryomyces hansenii was investigated at different pH and temperature values. At optimum starting pH (pH(o) = 5.5), T = 24 degrees C, and relatively low starting biomass levels (0.5-0.6 g(x)/L), 88% of xylose was utilized for xylitol production, the rest being preferentially fermented to ethanol (10%). Under these conditions, nearly 70% of initial carbon was recovered as xylitol, corresponding to final xylitol concentration of 91.9 g(P)/L, product yield on substrate of 0.81 g(P)/g(S), and maximum volumetric and specific productivities of 1.86 g(P)/L x h and 1.43 g(P)/g(x) x h, respectively. At higher and lower pH(o) values, respiration also became important, consuming up to 32% of xylose, while negligible amounts were utilized for cell growth (0.8-1.8%). The same approach extended to the effect of temperature on the metabolism of this yeast at pH(o) = 5.5 and higher biomass levels (1.4-3.0 g(x)/L) revealed that, at temperatures ranging from 32-37 degrees C, xylose was nearly completely consumed to produce xylitol, reaching a maximum volumetric productivity of 4.67 g(P)/L x h at 35 degrees C. Similarly, both respiration and ethanol fermentation became significant either at higher or at lower temperatures. Finally, to elucidate the kinetic mechanisms of both xylitol production and thermal inactivation of the system, the related thermodynamic parameters were estimated from the experimental data with the Arrhenius model: activation enthalpy and entropy were 57.7 kJ/mol and -0.152 kJ/mol x K for xylitol production and 187.3 kJ/mol and 0.054 kJ/mol x K for thermal inactivation, respectively.  相似文献   

3.
About 270 yeast isolates were screened for xylitol production using xylose as the sole carbon source. The best isolate, Debaryomyces hansenii UFV-170, released 5.84 g L(-1) xylitol from 10 g L(-1) xylose after 24 h, corresponding to a yield of xylitol on consumed substrate (Y(P/S)) of 0.54 g g(-1). This strain was cultivated batch-wise at variable starting concentrations of xylose (S(o)) and biomass (X(o)) and agitation intensity, in order to improve xylitol production and to evaluate, through simple carbon balances, the influence of these conditions on xylose metabolism. Under the best microaerobic conditions (S(o) = 53 g L(-1), X(o) = 1.4 g L(-1), 200 rpm), xylitol production reached 37.0 g L(-1), corresponding to xylitol volumetric productivity of 1.0 g L(-1)h(-1), specific productivity of 0.22 g g(-1)h(-1) and Y(P/S) = 0.76 g g(-1). Almost 83% of xylose was consumed for xylitol production, the rest being consumed for growth, while respiration was negligible. The new isolate appeared to be a promising alternative for industrial xylitol bioproduction.  相似文献   

4.
The effect of oxygenation on xylitol production by the yeast Debaryomyces hansenii has been investigated in this work using the liquors from corncob hydrolysis as the fermentation medium. The concentrations of consumed substrates (glucose, xylose, arabinose, acetate and oxygen) and formed products (xylitol, arabitol, ethanol, biomass and carbon dioxide) have been used, together with those previously obtained varying the hydrolysis technique, the level of adaptation of the microorganism, the sterilization procedure and the initial substrate and biomass concentrations, in carbon material balances to evaluate the percentages of xylose consumed by the yeast for the reduction to xylitol, alcohol fermentation, respiration and cell growth. The highest xylitol concentration (71 g/L) and volumetric productivity (1.5 g/L.h) were obtained semiaerobically using detoxified hydrolyzate produced by autohydrolysis-posthydrolysis, at starting levels of xylose (S(0)) and biomass (X(0)) of about 100 g/L and 12 g(DM)/L, respectively. No less than 80% xylose was addressed to xylitol production under these conditions. The experimental data collected in this work at variable oxygen levels allowed estimating a P/O ratio of 1.16 mol(ATP)/mol(O). The overall ATP requirements for biomass production and maintenance demonstrated to remarkably increase with X(0) and for S(0) >or= 130 g/L and to reach minimum values (1.9-2.1 mol(ATP)/C-mol(DM)) just under semiaerobic conditions favoring xylitol accumulation.  相似文献   

5.
The internal conductance to CO(2) transfer from intercellular spaces to chloroplasts poses a major limitation to photosynthesis, but few studies have investigated its temperature response. The aim of this study was to determine the temperature response of photosynthesis and internal conductance between 10 degrees C and 35 degrees C in seedlings of a deciduous forest tree species, Quercus canariensis. Internal conductance was estimated via simultaneous measurements of gas exchange and chlorophyll fluorescence ("variable J method"). Two of the required parameters, the intercellular photocompensation point (C(i)*) and rate of mitochondrial respiration in the light (R(d)), were estimated by the Laisk method. These were used to calculate the chloroplastic photocompensation point (Gamma*) in a simultaneous equation with g(i). An independent estimate of internal conductance was obtained by a novel curve-fitting method based on the curvature of the initial Rubisco-limited portion of an A/C(i) curve. The temperature responses of the rate of Rubisco carboxylation (V(cmax)) and the RuBP limited rate of electron transport (J(max)) were determined from chloroplastic CO(2) concentrations. The rate of net photosynthesis peaked at 24 degrees C. C(i)* was similar to reports for other species with a C(i)* of 39 micromol mol(-1) at 25 degrees C and an activation energy of 34 kJ mol(-1). Gamma* was very similar to the published temperature response for Spinacia oleracea from 20 degrees C to 35 degrees C, but was slightly greater at 10 degrees C and 15 degrees C. J(max) peaked at 30 degrees C, whereas V(cmax) did not reach a maximum between 10 degrees C and 35 degrees C. Activation energies were 49 kJ mol(-1) for V(cmax) and 100 kJ mol(-1) for J(max). Both methods showed that internal conductance doubled from 10 degrees C to 20 degrees C, and then was nearly temperature-independent from 20 degrees C to 35 degrees C. Hence, the temperature response of internal conductance could not be fitted to an Arrhenius function. The best fit to estimated g(i) was obtained with a three-parameter log normal function (R(2)=0.98), with a maximum g(i) of 0.19 mol m(-2) s(-1) at 29 degrees C.  相似文献   

6.
The thermodynamic stability of staphylococcal nuclease was studied against the variation of both temperature and pressure by utilizing (1)H NMR spectroscopy at 750 MHz in 20 mM Mes buffer containing 99.9 % (2)H(2)O, pH 5.3. Equilibrium fractions of folded and unfolded protein species were evaluated with the proton signals of two histidine residues as monitor in the pressure range of 30-3300 bar and in the temperature range of 1.5 degrees C-35 degrees C. From the multi-parameter fit of the experimental data to the Gibbs energy equation expressed as a simultaneous function of pressure and temperature, we determined the compressibility change (Deltabeta), the volume change at 1 bar (DeltaV degrees ) and the expansivity change (Deltaalpha) upon unfolding among other thermodynamic parameters: Deltabeta=0.02(+/-0.003) ml mol(-1) bar(-1); Deltaalpha=1.33(+/-0.2) ml mol(-1) K(-1); DeltaV degrees =-41.9(+/-6. 3) ml mol(-1) (at 24 degrees C); DeltaG degrees =13.18(+/-2) kJ mol(-1) (at 24 degrees C); DeltaC(p)=13.12(+/-2) kJ mol(-1) K(-1); DeltaS degrees =0.32(+/-0.05) kJ mol(-1) K(-1 )(at 24 degrees C). The result yields a three-dimensional free energy surface, i.e. the free energy-landscape of staphylococcal nuclease on the P-T plane. The significantly positive Deltabeta and Deltaalpha values suggest that, in the pressure-denatured state, staphylococcal nuclease forms a loosely packed and fluctuating structure. The slight but statistically significant difference between the unfolding transitions of the His8 and His124 environments is considered to reflect local fluctuations in the native state, leading to pre-melting of the His124 environment prior to the cooperative unfolding of the major part of the protein.  相似文献   

7.
The kinetic and thermodynamic properties of ascorbate oxidase (AO) activity and stability of a Cucurbita maxima extract were investigated. Activity tests performed at 25 degrees C using initial ascorbic acid concentration in the range 50-750 M allowed estimating the Michaelis constant for this substrate (Km = 126 microM) and the maximum initial rate of ascorbic acid oxidation (A0,max = 1.57 mM min-1). The main thermodynamic parameters of the enzyme reaction (DeltaH* = 10.3 kJ mol-1; DeltaG* = 87.2 kJ mol-1; DeltaS* = -258 J mol-1 K-1) were estimated through activity tests performed at 25-48 C. Within such a temperature range, no decrease in the initial reaction rate was detected. The long-term thermostability of the raw extract was then investigated by means of residual activity tests carried out at 10-70 degrees C, which allowed estimating the thermodynamic parameters of the irreversible enzyme inactivation as well (DeltaH*D = 51.7 kJ mol-1; DeltaG*D = 103 kJ mol-1; S*D = -160 J mol-1 K-1). Taking into account the specific rate of AO inactivation determined at different temperatures, we also estimated the enzyme half-life (1047 min at 10 degrees C and 21.2 min at 70 degrees C) and predicted the integral activity of a continuous system using this enzyme preparation. This work should be considered as a preliminary attempt to characterize the AO activity of a C. maxima extract before its concentration by liquid-liquid extraction techniques.  相似文献   

8.
The thermodynamics of the conversion of aqueous L-aspartic acid to fumaric acid and ammonia have been investigated using both heat conduction microcalorimetry and high-pressure liquid chromatography. The reaction was carried out in aqueous phosphate buffer over the pH range 7.25-7.43, the temperature range 13-43 degrees C, and at ionic strengths varying from 0.066 to 0.366 mol kg(-1). The following values have been found for the conversion of aqueous L-aspartateH- to fumarate2- and NH4+ at 25 degrees C and at zero ionic strength: K = (1.48 +/- 0.10) x 10(-3), DeltaG degrees = 16.15 +/- 0.16 kJ mol(-1), DeltaH degrees = 24.5 +/- 1.0 kJ mol(-1), and DeltaC(p) degrees = -147 +/- 100 J mol(-1) K(-1). Calculations have also been performed which give values of the apparent equilibrium constant for the conversion of L-aspartic acid to fumaric acid and ammonia as a function of temperature, pH and ionic strength.  相似文献   

9.
Escherichia coli strain PC09 (DeltaxylB, cAMP-independent CRP (crp*) mutant) expressing an NADPH-dependent xylose reductase from Candida boidinii (CbXR) was previously reported to produce xylitol from xylose while metabolizing glucose [Cirino et al. (2006) Biotechnol Bioeng 95(6): 1167-1176]. This study aims to understand the role of NADPH supply in xylitol yield and the contribution of key central carbon metabolism enzymes toward xylitol production. Studies in which the expression of CbXR or a xylose transporter was increased suggest that enzyme activity and xylose transport are not limiting xylitol production in PC09. A constraints-based stoichiometric metabolic network model was used to understand the roles of central carbon metabolism reactions and xylose transport energetics on the theoretical maximum molar xylitol yield (xylitol produced per glucose consumed), and xylitol yields (Y(RPG)) were measured from resting cell biotransformations with various PC09 derivative strains. For the case of xylose-proton symport, omitting the Zwf (glucose-6-phosphate dehydrogenase) or PntAB (membrane-bound transhydrogenase) reactions or TCA cycle activity from the model reduces the theoretical maximum yield from 9.2 to 8.8, 3.6, and 8.0 mol xylitol (mol glucose)(-1), respectively. Experimentally, deleting pgi (encoding phosphoglucose isomerase) from strain PC09 improves the yield from 3.4 to 4.0 mol xylitol (mol glucose)(-1), while deleting either or both E. coli transhydrogenases (sthA and pntA) has no significant effect on the measured yield. Deleting either zwf or sucC (TCA cycle) significantly reduces the yield from 3.4 to 2.0 and 2.3 mol xylitol (mol glucose)(-1), respectively. Expression of a xylose reductase with relaxed cofactor specificity increases the yield to 4.0. The large discrepancy between theoretical maximum and experimentally determined yield values suggests that biocatalysis is compromised by pathways competing for reducing equivalents and dissipating energy. The metabolic role of transhydrogenases during E. coli biocatalysis has remained largely unspecified. Our results demonstrate the importance of direct NADPH supply by NADP+-utilizing enzymes in central metabolism for driving heterologous NADPH-dependent reactions, and suggest that the pool of reduced cofactors available for biotransformation is not readily interchangeable via transhydrogenase.  相似文献   

10.
The effect of pH and temperature on the apparent association equilibrium constant (Ka) for the binding of the recombinant proteinase inhibitor eglin c (eglin c), of the soybean Bowman-Birk proteinase inhibitor (BBI) and of its chymotrypsin and trypsin inhibiting fragments (F-C and F-T, respectively) to Leu-proteinase, the leucine specific serine proteinase from spinach (Spinacia oleracea L.) leaves, has been investigated. On lowering the pH from 9.5 to 4.5, values of Ka (at 21 degrees C) for complex formation decrease thus reflecting the acidic pK-shift of the hystidyl catalytic residue from approximately 6.9, in the free Leu-proteinase, to approximately 5.1, in the enzyme: inhibitor adducts. At pH 8.0, values of the apparent thermodynamic parameters for the proteinase:inhibitor complex formation are: Leu-proteinase:eglin c-Ka = 2.2 x 10(11) M-1, delta G degree = -64 kJ/mol, delta H degree = +5.9 kJ/mol, and delta S degree = +240 kJ/molK; Leu-proteinase:BBI-Ka = 3.2 x 10(10) M-1, delta G degree = -59 kJ/mol, delta H degree = +8.8 kJ/mol, and delta S degree = +230 J/molK; and Leu-proteinase:F-C-Ka = 1.1 x 10(6) M-1, delta G degree = -34 kJ/mol, delta H degree = +18 J/mol, and delta S degree = +180 J/molK (values of Ka, delta G degree and delta S degree were obtained at 21.0 degrees C; values of delta H degree were temperature-independent over the range explored, i.e. between 10.0 degrees C and 40.0 degrees C).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
M S Matta  M E Andracki 《Biochemistry》1988,27(21):8000-8007
The specificity ratios kc/Km = k for subtilisin A catalyzed hydrolysis of five aryl esters of N-(methoxycarbonyl)-L-Phe (McPhe) were determined at pH 7.03 and its pD equivalent. The ratios are independent of the electronic properties of the leaving group substituent. Kinetic solvent isotope effects, Dk, increase from about 0.9 to 1.3 as leaving group ability decreases from p-nitrophenolate to p-methoxyphenolate. The k of N-(methoxycarbonyl)-L-phenylalanine p-nitrophenyl ester (NPE) with native enzyme exhibits a strong temperature dependence; delta H* = 87 +/- 3 kJ mol-1 and delta S* = 148 +/- 14 J K-1 mol-1 at 25 degrees C (H2O). The Dk with this substrate is 1.36 at 13.6 degrees C, declines to 0.89 at 25 degrees C, and then increases to 1.04 at 39.4 degrees C. Above neutral pH(D), with McPhe NPE as substrate, the dependence of k is for the dissociated form of a single base of pKapp = 7.38 +/- 0.03 in H2O and 7.67 +/- 0.03 in D2O. The pKapp values are apparently those of the uncomplexed native protein. By contrast, k of 3-phenylpropanoic acid (Prop) p-nitrophenyl ester exhibits a weaker temperature dependence; delta H* = 20 kJ mol-1 and delta S* = -90 J K-1 mol-1 (H2O) at 25 degrees C. The Dk are larger than those for McPhe NPE, decreasing from 1.99 at 20.5 degrees C to 1.74 at 46.1 degrees C. These results, combined with those of previous studies, are consistent with limitation of k by at least two processes.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
The 1,044 bp endo-1,4-β-xylanase gene of a hyperthermophilic Eubacterium, "Thermotoga petrophila RKU 1" (T. petrophila) was amplified, from the genomic DNA of donor bacterium, cloned and expressed in mesophilic host E. coli strain BL21 Codon plus. The extracellular target protein was purified by heat treatment followed by anion and cation exchange column chromatography. The purified enzyme appeared as a single band, corresponding to molecular mass of 40 kDa, upon SDS-PAGE. The pH and temperature profile showed that enzyme was maximally active at 6.0 and 95 °C, respectively against birchwood xylan as a substrate (2,600 U/mg). The enzyme also exhibited marked activity towards beech wood xylan (1,655 U/mg). However minor activity against CMC (61 U/mg) and β-Glucan barley (21 U/mg) was observed. No activity against Avicel, Starch, Laminarin and Whatman filter paper 42 was observed. The K(m), V(max) and K (cat) of the recombinant enzyme were found to be 3.5 mg ml(-1), 2778 μmol mg(-1)min(-1) and 2,137,346.15 s(-1), respectively against birchwood xylan as a substrate. The recombinant enzyme was found very stable and exhibited half life (t(?)) of 54.5 min even at temperature as high as 96 °C, with enthalpy of denaturation (ΔH*(D)), free energy of denaturation (ΔG*(D)) and entropy of denaturation (ΔS*(D)) of 513.23 kJ mol(-1), 104.42 kJ mol(-1) and 1.10 kJ mol(-1)K(-1), respectively at 96 °C. Further the enthalpy (ΔH*), Gibbs free energy (ΔG*) and entropy (ΔS*) for birchwood xylan hydrolysis by recombinant endo-1,4-β-xylanase were calculated at 95 °C as 62.45 kJ mol(-1), 46.18 kJ mol(-1) and 44.2 J mol(-1) K(-1), respectively.  相似文献   

13.
Kinetics for the breakdown of the trinuclear chromium acetate cluster, [Cr(3)O(OAc)(6)](+), with a series of monoprotic and diprotic ligands in weakly acidic aqueous media (pH approximately 4 or approximately 5) have been investigated spectrophotometrically at 40-60 degrees C. The results point to an ion-pair equilibrium as the first step followed by associative interchange mechanism forming the mononuclear product of the reaction. Pseudo-first-order rates were determined from absorbance data and associated activation parameters were calculated using the Eyring equation. Enthalpy and entropy terms of the reactions (e.g., histidine, DeltaH(double dagger) = 75 +/- 15 kJ mol(-1), DeltaS(double dagger) = -130 +/- 25 J K(-1) mol(-1); lactic acid, DeltaH(double dagger) = 66 +/- 13 kJ mol(-1), DeltaS(double dagger) = -155 +/- 30 J K(-1) mol(-1); glycine, DeltaH(double dagger) = 31 +/- 6 kJ mol(-1), DeltaS(double dagger) = -225 +/- 45 J K(-1) mol(-1)) are consistent with an associative interchange (I(a)) mechanism, and produce a linear isokinetic plot (slope = 50 degrees C). Rates and activation parameters are comparable to those of substitution reactions of the chromium(III) hexaaqua cation. Other ligands studied included malonic acid and the amino acid, aspartic acid. Observed rates are faster than water exchange rates, but typically slower than anion substitution rates, and indicate that trinuclear chromium(III) clusters are expected to be kinetically stable in neutral to slightly acidic conditions.  相似文献   

14.
Xylose reductase has been purified to apparent homogeneity from cell extracts of the fungus Cryptococcus flavus grown on D-xylose as carbon source. The enzyme, the first of its kind from the phylum Basidiomycota, is a functional dimer composed of identical subunits of 35.3 kDa mass and requires NADP(H) for activity. Steady-state kinetic parameters for the reaction, D-xylose + NADPH + H(+)<--> xylitol + NADP(+), have been obtained at pH 7.0 and 25 degrees C. The catalytic efficiency for reduction of D-xylose is 150 times that for oxidation of xylitol. This and the 3-fold tighter binding of NADPH than NADP(+) indicate that the enzyme is primed for unidirectional metabolic function in microbial physiology. Kinetic analysis of enzymic reduction of aldehyde substrates differing in hydrophobic and hydrogen bonding capabilities with binary enzyme-NADPH complex has been used to characterize the substrate-binding pocket of xylose reductase. Total transition state stabilization energy derived from bonding with non-reacting sugar hydroxyls is approximately 15 kJ/mol, with a major contribution of 5-8 kJ/mol made by interactions with the C-2(R) hydroxy group. The aldehyde binding site is approximately 1.2 times more hydrophobic than n-octanol and can accommodate linear alkyl chains of 相似文献   

15.
Escherichia coli W3110 was previously engineered to co-utilize glucose and xylose by replacing the wild-type crp gene with a crp* mutant encoding a cAMP-independent CRP variant (Cirino et al., 2006 [Cirino, P.C., Chin, J.W., Ingram, L.O., 2006. Engineering Escherichia coli for xylitol production from glucose-xylose mixtures. Biotechnol. Bioeng. 95, 1167-1176.]). Subsequent deletion of the xylB gene (encoding xylulokinase) and expression of xylose reductase from Candida boidinii (CbXR) resulted in a strain which produces xylitol from glucose-xylose mixtures. In this study we examine the contributions of the native E. coli xylose transporters (the d-xylose/proton symporter XylE and the d-xylose ABC transporter XylFGH) and CRP* to xylitol production in the presence of glucose and xylose. The final batch xylitol titer with strain PC09 (Delta xylB and crp*) is reduced by 40% upon deletion of xylG and by 60% upon deletion of both xyl transporters. Xylitol production by the wild-type strain (W3110) expressing CbXR is not reduced when xylE and xylG are deleted, demonstrating tight regulation of the xylose transporters by CRP and revealing significant secondary xylose transport. Finally, plasmid expression of XylE or XylFGH with CbXR in PC07 (Delta xylB and wild-type crp) growing on glucose results in xylitol titers similar to that achieved with PC09 and provides an alternative strategy to the use of CRP*.  相似文献   

16.
AIMS: To investigate the production of xylitol by the yeast Candida guilliermondii FTI 20037, in a bioreactor, from rice straw hemicellulosic hydrolysate with a high xylose concentration. METHODS AND RESULTS: Batch fermentation was carried out with rice straw hemicellulosic hydrolysate containing about 85 g xylose l(-1), in a stirred-tank bioreactor at 30 degrees C, under aeration of 1.3 vvm (volume of air per volume of medium per min) and different stirring rates (200, 300 and 500 rev min(-1)). The bioconversion of xylose into xylitol by the yeast depended on the stirring rate, the maximum xylitol yield (YP/S = 0.84 g g(-1)) being achieved at 300 rev min-1, with no need to pretreat the hydrolysate for purification. CONCLUSIONS: To determine the most adequate oxygen transfer rate is fundamental to improving the xylose-to-xylitol bioconversion by C. guilliermondii. SIGNIFICANCE AND IMPACT OF THE STUDY: For the microbial production of xylitol to be economically viable, the initial concentration of xylose in the lignocellulosic hydrolysate should be as high as possible, as with high substrate concentrations it is possible to increase the final product concentration. Nevertheless, there are few reports on the use of high xylose concentrations. Considering a process in bioreactor, from rice straw hemicellulosic hydrolysate, this is an innovator work.  相似文献   

17.
Production of xylitol from D-xylose by recombinant Lactococcus lactis   总被引:1,自引:0,他引:1  
The D-xylose reductase from Pichia stipitis CBS 5773 and the xylose transporter from Lactobacillus brevis ATCC 8287 were expressed in active form in Lactococcus lactis NZ9800. Xylitol production was investigated using non-growing recombinant cells in high cell-density under microaerobic conditions in the presence of xylose and glucose. Besides xylose, the recombinant strain with xylose reductase activity reduced l-arabinose and D-ribose in significant extent to the corresponding pentitols. The ratio of xylitol produced per glucose consumed was almost 10-fold higher under glucose limitation than the ratio in the presence of excess initial glucose. The co-expression of the xylose transporter with the xylose reductase did not increase the efficiency of xylitol production appreciably when compared to the strain in which only the xylose reductase gene was expressed. A fed-batch experiment with high initial xylose concentration (160 gl(-1)) under glucose limitation was carried out using the strain co-expressing xylose reductase and xylose transporter genes. The xylitol yield from xylose was 1.0 mol mol(-1) and the ratio of xylitol produced per glucose consumed was 2.5 mol mol(-1). The volumetric productivity was 2.72 gl(-1)h(-1) at 20 h. Of the xylose initially present, 34% was consumed. Analysis of the fermentation metabolites revealed a shift from homolactic to mixed acid fermentation at early stages of the experiment.  相似文献   

18.
The kinetics of cyanide binding to chloroperoxidase were studied using a high-pressure stopped-flow technique at 25 degrees C and pH 4.7 in a pressure range from 1 to 1000 bar. The activation volume change for the association reaction is delta V not equal to + = -2.5 +/- 0.5 ml/mol. The total reaction volume change, determined from the pressure dependence of the equilibrium constant, is delta V degrees = -17.8 +/- 1.3 ml/mol. The effect of temperature was studied at 1 bar yielding delta H not equal to + = 29 +/- 1 kJ/mol, delta S not equal to + = -58 +/- 4 J/mol per K. Equilibrium studies give delta H degrees = -41 +/- 3 kJ/mol and delta S degrees = -59 +/- 10 J/mol per K. Possible contributions to the binding process are discussed: changes in spin state, bond formation and conformation changes in the protein. An activation volume analog of the Hammond postulate is considered.  相似文献   

19.
Arthrospira platensis was cultivated photoautotrophically at 6.0 klux light intensity in 5.0-L open tanks, using a mineral medium containing urea as nitrogen source. Fed-batch experiments were performed at constant flowrate. A central composite factorial design combined to response surface methodology (RSM) was utilized to determine the relationship between the selected response variables (cell concentration after 10 days, X(m), cell productivity, P(X), and nitrogen-to-cell conversion factor, Y(X/N)) and codified values of the independent variables (pH, temperature, T, and urea flowrate, K). By applying the quadratic regression analysis, the equations describing the behaviors of these responses as simultaneous functions of the selected independent variables were determined, and the conditions for X(m) and P(X) optimization were estimated (pH 9.5, T = 29 degrees C, and K = 0.551 mM/day). The experimental data obtained under these conditions (X(m) = 749 mg/L; P(X) = 69.9 mg/L.day) were very close to the estimated ones (X(m) = 721 mg/L; P(X) = 67.1 mg/L.day). Additional cultivations were carried out under the above best conditions of pH control and urea flowrate at variable temperature. Consistently with the results of RSM, the best growth temperature was 29 degrees C. The maximum specific growth rates at different temperatures were used to estimate the thermodynamic parameters of growth (DeltaH* = 59.3 kJ/mol; DeltaS* = -0.147 kJ/mol.K; DeltaG* = 103 kJ/mol) and its thermal inactivation (DeltaH(D) (o) = 72.0 kJ/mol; DeltaS(D) (o) = 0.144 kJ/mol.K; DeltaG(D) (o) = 29.1 kJ/mol).  相似文献   

20.
The binding of UDP-N-acetylglucosamine (UDPNAG) to the enzyme UDP-N-acetylglucosamine enolpyruvyl transferase (MurA) was studied in the absence and presence of the antibiotic fosfomycin by isothermal titration calorimetry. Fosfomycin binds covalently to MurA in the presence of UDPNAG and also in its absence as demonstrated by MALDI mass spectrometry. The covalent attachment of fosfomycin affects the thermodynamic parameters of UDPNAG binding significantly: In the absence of fosfomycin the binding of UDPNAG is enthalpically driven (DeltaH = -35.5 kJ mol(-1) at 15 degrees C) and opposed by an unfavorable entropy change (DeltaS = -25 J mol(-1) K(-1)). In the presence of covalently attached fosfomycin the binding of UDPNAG is entropically driven (DeltaS = 187 J mol(-1)K(-1) at 15 degrees C) and associated with unfavorable changes in enthalpy (DeltaH = 28.8 kJ mol(-1)). Heat capacities for UDPNAG binding in the absence or presence of fosfomycin were -1.87 and -2.74 kJ mol(-1) K(-1), respectively, indicating that most ( approximately 70%) of the conformational changes take place upon formation of the UDPNAG-MurA binary complex. The major contribution to the heat capacity of ligand binding is thought to be due to changes in the solvent-accessible surface area. However, associated conformational changes, if any, also contribute to the experimentally measured magnitude of the heat capacity. The changes in solvent-accessible surface area were calculated from available 3D structures, yielding a DeltaC(p) of -1.3 kJ mol(-1) K(-1); i.e., the experimentally determined heat capacity exceeds the calculated one. This implies that other thermodynamic factors exert a large influence on the heat capacity of protein-ligand interactions.  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