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.
Studies of the photochemical reactivity of pterin (= 2-aminopteridin-4(3H)-one; PT) in acidic (pH 5.0-6.0) and alkaline (pH 10.2-10.8) aqueous solutions have been performed. The photochemical reactions were followed by UV/VIS spectrophotometry, thin layer chromatography (TLC), high-performance liquid chromatography (HPLC), and an enzymatic method for H2O2 determination. PT is not light-sensitive in the absence of molecular oxygen, but it undergoes photooxidation in the presence of O2, yielding several nonpteridinic products. The quantum yields for PT disappearance were found to be 8.2 (+/-0.6) x 10(-4) and 1.2 (+/-0.2) x 10(-3) in acidic and alkaline media, respectively. H2O2 was detected and quantified in irradiated solutions of PT; and its importance from a biomedical point of view is discussed. The rate constant of the chemical reaction between singlet oxygen ((1)O2) and PT was determined to be 2.5 (+/-0.2) x 10(5) l mol(-1) s(-1) in alkaline medium, and the role of (1)O2 in the photooxidation of pterin was evaluated.  相似文献   

2.
To better understand substrate recognition and catalysis by RNase III, we examined steady-state and pre-steady-state reaction kinetics, and changes in intrinsic enzyme fluorescence. The multiple turnover cleavage of a model RNA substrate shows a pre-steady-state burst of product formation followed by a slower phase, indicating that the steady-state reaction rate is not limited by substrate cleavage. RNase III catalyzed hydrolysis is slower at low pH, permitting the use of pre-steady-state kinetics to measure the dissociation constant for formation of the enzyme-substrate complex (K(d)=5.4(+/-0.6) nM), and the rate constant for phosphodiester bond cleavage (k(c)=1.160(+/-0.001) min(-1), pH 5.4). Isotope incorporation analysis shows that a single solvent oxygen atom is incorporated into the 5' phosphate of the RNA product, which demonstrates that the cleavage step is irreversible. Analysis of the pH dependence of the single turnover rate constant, k(c), fits best to a model for two or more titratable groups with pK(a) of ca 5.6, suggesting a role for conserved acidic residues in catalysis. Additionally, we find that k(c) is dependent on the pK(a) value of the hydrated divalent metal ion included in the reaction, providing evidence for participation of a metal ion hydroxide in catalysis, potentially in developing the nucleophile for the hydrolysis reaction. In order to assess whether conformational changes also contribute to the enzyme mechanism, we monitored intrinsic tryptophan fluorescence. During a single round of binding and cleavage by the enzyme we detect a biphasic change in fluorescence. The rate of the initial increase in fluorescence was dependent on substrate concentration yielding a second-order rate constant of 1.0(+/-0.1)x10(8) M(-1) s(-1), while the rate constant of the second phase was concentration independent (6.4(+/-0.8) s(-1); pH 7.3). These data, together with the unique dependence of each phase on divalent metal ion identity and pH, support the hypothesis that the two fluorescence transitions, which we attribute to conformational changes, correlate with substrate binding and catalysis.  相似文献   

3.
Martin BL  Jurado LA  Hengge AC 《Biochemistry》1999,38(11):3386-3392
Activation of calcineurin by Mn2+ and Mg2+ was compared using a heavy atom isotope analogue of the substrate p-nitrophenyl phosphate (pNPP). Heavy atom isotope effects were measured for Mg2+ activation and compared to published results of the isotope effects with Mn2+ as the activating metal. Isotope effects were measured for the kinetic parameter Vmax/Km at the nonbridging oxygen atoms [18(V/K)nonbridge]; at the position of bond cleavage in the bridging oxygen atom [18(V/K)bridge]; and at the nitrogen atom in the nitrophenol leaving group [15(V/K)]. The isotope effects increased in magnitude upon changing from an optimal pH to a nonoptimal pH; the 18(V/K)bridge effect increased from 1.0154 (+/-0.0007) to 1.0198 (+/-0.0002), and the 15(V/K) effect increased from 1.0018 (+/-0. 0002) to 1.0021 (+/-0.0003). The value for 18(V/K)nonbridge is 0. 9910 (+/-0.0003) at pH 7.0. As with Mn2+, the 18(V/K)nonbridge isotope effect indicated that the dianion was the substrate for catalysis, and that a dissociative transition state was operative for the phosphoryl transfer. Comparison to results for Mn2+ activation suggested that chemistry was more rate-limiting with Mg2+ than with Mn2+. Changing the activating metal concentration showed opposite trends with increasing Mg2+ increasing the commitment factor and seemingly making the chemistry less rate-limiting. The influence of viscosity was evaluated as well to gauge the role of chemistry. The activation of calcineurin-catalyzed hydrolysis of pNPP1 by Mg2+ or Mn2+ at pH 7.0 was compared in the presence of viscogens, glycerol and poly(ethylene glycol). Increasing glycerol caused different effects with the two activators. With Mn2+ as the activator, calcineurin activity showed a normal response with kcat and kcat/Km decreasing with viscosity. There was an inverse response with Mg2+ as the activator as values of kcat/Km increased with viscosity. From values of the normalized kcat/Km with Mn2+, the chemistry was found to be partially rate-limiting, consistent with previous heavy atom isotope studies (22). The effect observed for Mg2+ seems consistent with a change in the rate-limiting step for the two different metals at pH 7.0.  相似文献   

4.
A D Hall  A Williams 《Biochemistry》1986,25(17):4784-4790
Values of kcat and Km have been measured for the Escherichia coli alkaline phosphatase catalyzed hydrolysis of 18 aryl and 12 alkyl monophosphate esters at pH 8.00 and 25 degrees C. A Br?nsted plot of log (kcat/Km) (M-1 s-1) vs. the pK of the leaving hydroxyl group exhibits two regression lines: log (kcat/Km) = -0.19 (+/- 0.02) pKArOH + 8.14 (+/- 0.15) log (kcat/Km) = -0.19 (+/- 0.01) pKROH + 5.89 (+/- 0.17) Alkyl phosphates with aryl or large lipophilic side chains are not correlated by the above equations and occupy positions intermediate between the two lines. The observed change in effective charge on the leaving oxygen of the ester (-0.2) is very small, consistent with substantial electrophilic participation of the enzyme with this atom. Cyclohexylammonium ion is a noncompetitive inhibitor against 4-nitrophenyl phosphate substrate at pH 8.00, and neutral phenol is a competitive inhibitor (Ki = 82.6 mM); these data and the 100-fold larger reactivity of aryl over alkyl esters are consistent with the existence of a lipophilic binding site for the leaving group of the substrate. The absence of a major steric effect in kcat/Km for substituted aryl esters confirms that the leaving group in the enzyme--substrate complex points away from the surface of the enzyme. Arguments are advanced to exclude a dissociative mechanism (involving a metaphosphate ion) for the enzyme-catalyzed substitution at phosphorus.  相似文献   

5.
Elongation rate and heat produced by Avena coleoptile segments suspended in sucrose buffer solutions were measured at pH values from 3.5 to 8.5. The caloric efficiency of elongation (CEE) was defined as the ratio of the rate of elongation to the rate of heat production. Elongation and CEE were greatest at intermediate pH values, but heat production (about 1 cal/g.hr) was insensitive to pH within the limits of experimental error (+/-20%). Quantitative agreement was found between the results of previous respiration studies and the rate of heat production in an aerobic atmosphere, which indicates that oxidative metabolism accounts for essentially all energy changes in the cell, so matter flow is a significant component of the bioenergetics of cell function. Indole-3-acetic acid up to 1 mm, produced about a 10-fold increase in elongation rate, a 5-fold increase of the CEE, and a 25% increase in heat production. Above this concentration, sharp drops in both elongation and heat production occurred, without altering the CEE at pH 6.5, but greatly decreasing the CEE at pH 4.5. Elongation and CEE showed marked decreases after 4 hours in an anaerobic atmosphere, but heat production did not exhibit a proportional decrease. These studies indicate that rate of cell elongation in the presence and absence of auxin is not directly proportional to the overall metabolism of the cell.  相似文献   

6.
In 18 patients with gastrointestinal manifestations of digoxin toxicity the mean serum digoxin concentration (+/- SEM) was 3.16 micrograms/l (+/- 0.25), the calcium to potassium ratio 0.31 (+/- 0.01), and the mean arterial pH 7.406 (+/- 0.017). In contrast 19 patients with digoxin induced automaticity had a mean serum digoxin concentration of 1.24 micrograms/l (+/- 0.15; p less than 0.001), a calcium to potassium ratio of 0.38 (+/- 0.01; p less than 0.01), and an arterial pH of 7.498 (+/- 0.008; p less than 0.001). Eight out of 13 patients with digoxin induced cardiotoxicity had serum concentrations of the drug within the therapeutic range (0.8-2.0 micrograms/l). The calcium to potassium ratio, however, was lower than in the patients with automaticity (0.31 +/- 0.02; p less than 0.01) and the arterial pH was 7.370 (+/- 0.033; p less than 0.05). Serum magnesium concentrations were similar in all groups. In this study patients with digoxin induced gastrointestinal symptoms had high serum concentrations of the drug, whereas those with drug induced automaticity had therapeutic concentrations. This second group, however, was identified by their higher calcium to potassium ratios and higher pH values.  相似文献   

7.
Aliphatic thiols are effective as redox buffers for folding non-native disulfide-containing proteins into their native state at high pH values (8.0-8.5) but not at neutral pH values (6-7.5). In developing more efficient and flexible redox buffers, a series of aromatic thiols was analyzed for its ability to fold scrambled ribonuclease A (sRNase A). At equivalent pH values, the aromatic thiols folded sRNase A 10-23 times faster at pH 6.0, 7-12 times faster at pH 7.0, and 5-8 times faster at pH 7.7 than the standard aliphatic thiol glutathione. Similar correlations between thiol pK(a) values and folding rates at each pH value suggest that the apparent folding rate constants (k(app)) are a function of the redox buffer properties (pH, thiol pK(a) and [RSH]). Fitting the observed data to a three-variable model (logk(app)=-4.216(+/-0.030)+0.5816(+/-0.0036)pH-0.233(+/-0.004)pK(a)+log(1-e(-0.98(+/-0.02)[RSH]))) gave good statistics: r2=0.915, s=0.10.  相似文献   

8.
The oxygen consumption rate (VO(2)) of Biomphalaria glabrata populations, using polarometric and manometric methods, when plotted against dried body mass as logarithmic co-ordinates, respectively, fell on a regression line with a slope between 0.933 and 1.02. The slope of the regression line for non-infected Schistosoma mansoni populations was found to be 1.04 with no differences in the VO(2) between infected and non-infected snails. The VO(2) of CO-treated snails was the same as for the control snails. The VO(2) of starved snails declined after 3 days and was half the original value after 10 days starvation at 27 degrees C. The P(50) value for snail haemolymph containing haemoglobin suspended in a Tris-HCl buffer was 5.57(+/-0.73)mmHg at a pH of 7.51 and 25 degrees C. For Sephadex-75 cleaned haemolymph the P(50) value was 1.72(+/-0.07)mmHg at 25 degrees C and pH 7.51. Snails exposed to oxygen fs and to choices of different oxygen concentrations in water did not exclusively prefer high (130mmHg), low (15mmHg), or normal (80mmHg) oxygen tensions. The oxygen consumption rate of 782 cercariae at 27 degrees C was measured as 0.0092 microl O(2)/h per single cercaria. The results, when compared with the data in the literature [Z. Vergl. Physiol. 46 (1963) 467;; S. A. J. Zool. 14 (1979) 202], indicate that the mantle cavity gas bubble plays an insignificant or no role at all when pulmonate snails are kept in water with high partial pressures of oxygen and at low temperatures.  相似文献   

9.
The microscopic morphology, that is, total hyphal length and total number of tips, has been characterized during batch cultivations of Aspergillus oryzae. The specific growth rate estimated by measuring the total hyphal length (mu(h)) corresponds well with the specific growth rate estimated from dry weight measurements during cultures grown as free hyphal elements. The average tip extension rate can be described with a saturation type kinetics with respect to the average total hyphal length, and the branching frequency is closely related to the total hyphal length. For the applied strain of A. oryzae, pellet formation occurs by coagulation of spores. The agglomeration process is pH dependent and pellets are formed at pH values higher than 5, whereas low pH (<3.5) results in growth as freely dispersed hyphal elements. The maximum specific growth rate has a broad pH optimum between 3 and 7, whereas the alpha-amylase production has a sharper maximum at about pH 6. During batch cultivation with pellets the growth is described well by the cube-root law when pellet fragmentation can be neglected. The kinetic parameter k in the cube-root law is derived from the growth kinetics with no mass transfer limitation, k = mu(h)/3. Based on an oxygen balance, the active growth layer in the pellet is estimated to be 200 to 325 mum and, consequently, up to 50% of the biomass is limited by oxygen for large pellets. Ethanol production (up to 1 g L(-1)) was observed during batch cultivations with pellets, suggesting that ethanol is produced in the oxygen limited part of the biomass. A constitutive, low alpha-amylase production was observed at high glucose concentration. The specific alpha-amylase production was significantly higher for filamentous growth than for pellets and oxygen appears to be necessary for production of alpha-amylase. (c) 1996 John Wiley & Sons, Inc.  相似文献   

10.
Routine metabolic rate (RMR) was measured in fasting southern bluefin tuna, Thunnus maccoyii, the largest tuna species studied so far (body mass=19.6 kg (+/-1.9 SE)). Mean mass-specific RMR was 460 mg kg(-1) h(-1) (+/-34.9) at a mean water temperature of 19 degrees C. When evaluated southern bluefin tuna standard metabolic rate (SMR) is added to published values of other tuna species, there is a strong allometeric relationship with body mass (423 M(0.86), R(2)=0.97). This demonstrates that tuna interspecific SMR scale with respect to body mass similar to that of other active teleosts, but is approximately 4-fold higher. However, RMR (not SMR) is most appropriate in ram-ventilating species that are physiologically unable to achieve complete rest. Respiration was measured in a large (250,000 l) flexible polypropylene respirometer (mesocosm respirometer) that was deployed within a marine-farm sea cage for 29 days. Fasted fish were maintained within the respirometer up to 42 h while dissolved oxygen dropped by 0.056 (+/-0.004) mg l(-1) h(-1). Fish showed no obvious signs of stress. They swam at 1.1 (+/-0.1) fork lengths per second and several fed within the respirometer immediately after measurements.  相似文献   

11.
Routine metabolic rate (RMR) was measured in fasting southern bluefin tuna, Thunnus maccoyii, the largest tuna species studied so far (body mass=19.6 kg (+/-1.9 SE)). Mean mass-specific RMR was 460 mg kg(-1) h(-1) (+/-34.9) at a mean water temperature of 19 degrees C. When evaluated southern bluefin tuna standard metabolic rate (SMR) is added to published values of other tuna species, there is a strong allometeric relationship with body mass (423 M(0.86), R(2)=0.97). This demonstrates that tuna interspecific SMR scale with respect to body mass similar to that of other active teleosts, but is approximately 4-fold higher. However, RMR (not SMR) is most appropriate in ram-ventilating species that are physiologically unable to achieve complete rest. Respiration was measured in a large (250,000 l) flexible polypropylene respirometer (mesocosm respirometer) that was deployed within a marine-farm sea cage for 29 days. Fasted fish were maintained within the respirometer up to 42 h while dissolved oxygen dropped by 0.056 (+/-0.004) mg l(-1) h(-1). Fish showed no obvious signs of stress. They swam at 1.1 (+/-0.1) fork lengths per second and several fed within the respirometer immediately after measurements.  相似文献   

12.
SummarySelf-directing optimization was successfully employed to determine the optimal combination of engineering parameters, viz., pH, aeration rate and agitation rate, for extracellular ribonuclease production by Aspergillus niger SA-13-20 in a batch bioreactor. Maximal RNase production of 5.38 IU ml–1 was obtained at controlled pH of 2.33, aeration rate of 1.67 v/v/m and agitation rate of 850 rev/min. The effect of oxygen on the fermentation was also investigated. With increase in volumetric oxygen transfer coefficients (KLa), cell growth and RNase production first increased and then decreased. RNase production was further increased to 7.10 IU ml–1 and the fermentation time was shortened from 96 to 72 h by controlling dissolved oxygen concentration at 10% saturation by aerating oxygen after about 28 h of fermentation under the above optimal condition. The kinetic model showed that RNase production by A. niger SA-13-20 was growth-associated.  相似文献   

13.
Depending on solution conditions, beta-lactoglobulin can exist in one of its six pH-dependent structural states. We have characterized the acid and basic-induced conformational transitions between these structural states over the pH range of pH 1 to pH 13. To this end, we have employed high-precision ultrasonic and densimetric measurements coupled with fluorescence and CD spectroscopic data. Our combined spectroscopic and volumetric results have revealed five pH-induced transitions of beta-lactoglobulin between pH 1 and pH 13. The first transition starts at pH 2 and is not completed even at pH 1, our lowest experimental pH. This transition is followed by the dimer-to-monomer transition of beta-lactoglobulin between pH 2.5 and pH 4. The dimer-to-monomer transition is accompanied by decreases in volume, v degrees (-0.008(+/-0.003) cm3 x g(-1)), and adiabatic compressibility, k degrees (S) (-(0.7(+/-0.4))x10(-6) cm3 x g(-1) x bar(-1)). We interpret the observed changes in volume and compressibility associated with the dimer-to-monomer transition of beta-lactoglobulin, in conjunction with X-ray crystallographic data, as suggesting a 7 % increase in protein hydration, with the hydration changes being localized in the area of contact between the two monomeric subunits. The so-called N-to-Q transition of beta-lactoglobulin occurs between pH 4.5 and pH 6 and is accompanied by increases in volume, v degrees (0.004(+/-0.003) cm3 x g(-1)), and compressibility, k degrees (S) ((0.7(+/-0.4))x10(-6) cm3 x g(-1) x bar(-1)). The Tanford transition of beta-lactoglobulin is centered at pH 7.5 and is accompanied by a decrease in volume, v degrees (-0.006(+/-0.003) cm3 x g(-1)), and an increase in compressibility, k degrees (S) ((1.5(+/-0.5))x10(-6) cm3 x g(-1) x bar(-1)). Based on these volumetric results, we propose that the Tanford transition is accompanied by a 5 to 10 % increase in the protein hydration and a loosening of the interior packing of beta-lactoglobulin as reflected in a 12 % increase in its intrinsic compressibility. Finally, above pH 9, the protein undergoes irreversible base-induced unfolding which is accompanied by decreases in v degrees (-0.014(+/-0.003) cm3 x g(-1)) and k degrees (S) (-(7.0(+/-0.5))x10(-6) cm3 x g(-1) x bar(-1)). Combining these results with our CD spectroscopic data, we propose that, in the base-induced unfolded state of beta-lactoglobulin, only 80 % of the surface area of the fully unfolded conformation is exposed to the solvent. Thus, in so far as solvent exposure is concerned, the base-induced unfolded states of beta-lactoglobulin retains some order, with 20 % of its amino acid residues remaining solvent inaccessible.  相似文献   

14.
Two types of hydroponic bioreactors were used to investigate the mechanisnistic changes during phytoremediation of perchlorate under different root-zone conditions. The bioreactors included: (1) an aerobic ebb-and-flow system planted with six willow trees, and (2) individual willow trees grown in sealed root-zone bioreactors. Rhizosphere probes were used to monitor for the first time during phytoremediation of perchlorate, diurnal swings in oxidation-reduction potential (E(H)), dissolved oxygen (DO), and pH. Radiolabeled (36Cl-labeled) perchlorate was used as a tracer in a subset of the sealed bioreactor experiments to quantify the contribution of phytodegradation and rhizodegradation mechanisms. Rhizodegradation accounted for the removal of 96.1 +/- 4.5% (+/-95% CI) of the initial perchlorate dose in experiments conducted in sealed hydroponic bioreactors with low DO and little or no nitrate N. Meanwhile, the contribution of rhizodegradation decreased to 76 +/- 14% (+/-95% CI) when nitrate (a competing terminal electron acceptor) was provided as the nitrogen source. Slower rates of phytoremediation by uptake and phytodegradation were observed under high nitrate concentrations and aerobic conditions, which allowed perchlorate to persist in solution and resulted in a higher fraction uptake by the plant. Specifically, the rate of removal of perchlorate from bulk solution ranged from 5.4 +/- 0.54 to 37.1 +/- 2.25 mg/L/d (+/-SE) in the absence of nitrate to 1.78 +/- 0.27 to 0.46 +/- 0.02 mg/L/d (+/-SE) at high nitrate concentration. The results of this study indicate that the root-zone environment of plants can be manipulated to optimize rhizodegradation and to minimize undesirable processes such as uptake, temporal phytoaccumulation, and slow phytodegradation during phytoremediation of perchlorate. Rhizodegradation is desired because contaminants resident in plant tissue may remain an ecological risk until completely phytodegraded.  相似文献   

15.
溶氧及pH对产朊假丝酵母分批发酵生产谷胱甘肽的影响   总被引:16,自引:0,他引:16  
在7 L发酵罐中研究了溶氧和pH对产朊假丝酵母分批发酵生产谷胱甘肽的影响。结果表明,当葡萄糖浓度为30 g/L且通气量控制在5 L/min时,搅拌转速达到300 r/min即可满足细胞生长和谷胱甘肽合成对溶解氧的需求。不同pH控制方式对谷胱甘肽分批发酵的影响有较大差异。不控制pH时,细胞干重和谷胱甘肽产量比控制pH为55的发酵分别低27%和95%,且有50%的谷胱甘肽向胞外渗漏。研究了将pH控制在4.0、4.5、5.0、5.5、6.0和6.5的谷胱甘肽分批发酵过程,发现在pH 5.5时谷胱甘肽总产量最高。用前期研究建立的动力学模型模拟了不同pH (4.0~6.5)下的分批发酵过程,并从动力学角度解释了pH对细胞生长和谷胱甘肽合成的影响。  相似文献   

16.
The reduction of horse heart ferricytochrome c by the tryptic fragment of bovine liver cytochrome b5 and its dimethyl ester heme (DME)-substituted derivative has been studied as a function of ionic strength, pH, and temperature under solution conditions where the reaction is bimolecular. The rate constant for ferricytochrome c reduction by native ferrocytochrome b5 is 1.8 (+/- 0.2) x 10(7) M-1 s-1 (25 degrees C) with delta H++ = 7.5 (+/- 0.2) kcal/mol and delta S++ = -0.3 (+/- 0.6) eu (pH 7.0, I = 0.348 M). Under the same solution conditions, the reduction of ferricytochrome c by DME-ferrocytochrome b5 proceeds with a rate constant of 1.7 (+/- 0.1) x 10(7) M-1 s-1 with delta H++ = 7.9 (+/- 0.4) kcal/mol and delta S++ = 1 (+/- 1) eu. The rate constants for both reactions are strongly dependent on ionic strength. A detailed electrostatic analysis of the proteins has been performed. Two relatively simple Brownian dynamics simulation models predict rate constants for the reaction between the two native proteins that demonstrate a dependence on ionic strength similar to that observed experimentally. In one of these models, the proteins are treated as spheres with reactive surface patches that are defined by a 5 degrees cone generated about the dipole vector calculated for each protein and aligned with the presumed electron-transfer site near the partially exposed heme edge. The second model replaces the reactive patch assumption with an exponential distance dependence for the probability of reaction that permits estimation of a value for the distance-dependence factor alpha. Calculations with this latter model in combination with the aligned dipole assumption provide a reasonable approximation to the observed ionic strength dependence for the reaction and are consistent with a value of alpha = 1.2 A-1.  相似文献   

17.
Ca(2+) is essential for in vitro activity of Erwinia chrysanthemi pectate lyase C (PelC). Crystallographic analyses of 11 PelC-Ca(2+) complexes, formed at pH 4.5, 9.5, and 11.2 under varying Ca(2+) concentrations, have been solved and refined at a resolution of 2.2 A. The Ca(2+) site represents a new motif for Ca(2+), consisting primarily of beta-turns and beta-strands. The principal differences between PelC and the PelC-Ca(2+) structures at all pH values are the side-chain conformations of Asp-129 and Glu-166 as well as the occupancies of four water molecules. According to calculations of pK(a) values, the presence of Ca(2+) and associated structural changes lower the pK(a) of Arg-218, the amino acid responsible for proton abstraction during catalysis. The Ca(2+) affinity for PelC is weak, as the K(d) was estimated to be 0.132 (+/-0.004) mm at pH 9.5, 1.09 (+/-0.29) mm at pH 11.2, and 5.84 (+/-0.41) mm at pH 4.5 from x-ray diffraction studies and 0.133 (+/-0.045) mm at pH 9.5 from intrinsic tryptophan fluorescence measurements. Given the pH dependence of Ca(2+) affinity, PelC activity at pH 4.5 has been reexamined. At saturating Ca(2+) concentrations, PelC activity increases 10-fold at pH 4.5 but is less than 1% of maximal activity at pH 9.5. Taken together, the studies suggest that the primary Ca(2+) ion in PelC has multiple functions.  相似文献   

18.
Temperature jumps on mixtures of hemoglobin and pH indicators give rise to relaxation signals in the microsecond range. The pH and concentration dependences of the reciprocal relaxation time, 1/tau, may be rationalized on the basis of a reaction scheme in which a slow isomerization process in the protein moiety is coupled to a rapid co-operative ionization of two protons. At 11 degrees C the rate constants of the isomerization are kr = 4.2(+/- 1.8) x 10(4) s-1 and kf = 1.3(+/- 0.1) x 10(4) s-1 in deoxyhemoglobin; in carbonmonoxyhemoglobin they are kr = 3.9(+/- 1.3) x 10(4) s-1 and kf = 5.3(+/- 1.8) x 10(3) s-1. The pKa values of the coupled ionizing groups are 5.3 in deoxy- and 6.0 in carbonmonoxyhemoglobin. Modification of the CysF9(93) beta sulfhydryl group with iodoacetamide abolishes the pH dependence of 1/tau, suggesting that this sulfhydryl is involved in the isomerization process. Consideration of the X-ray structure of oxyhemoglobin allows a structural interpretation of the results. It is concluded that the isomerization may be important for the physiological function of hemoglobin, but that it is not identical with the quaternary structure transition.  相似文献   

19.
A macromodel for outdoor algal mass production   总被引:2,自引:0,他引:2  
A model describing growth of an outdoor algal (Spir-ulina platensis)culture was developed. The model can simulate biomass production, pH, growth rate, oxygen evolution, and CO(2) fixation rate. It was calibrated and validated against experimental data obtained by a novel automatic data logger/controller instrumentation which can number most vital parameters of the culture including on line estimation of oxygen production rate (OPR). The importance of understanding light distribution through the pond and its effects on the photosynthesis and respiration processes are emphasized. A maximum yield of about 38g day(-1) m(-2) under optimal conditions is predicted. The present model can also be a useful tool for optimization of algal mass production sites.  相似文献   

20.
中华绒螯蟹亲蟹的饥饿代谢研究   总被引:7,自引:0,他引:7  
1998年 10~ 12月 ,对体重为 5 4 .5 9(± 2 .37) g的亲蟹在 2 0 (± 0 .5 )℃温度条件下进行饥饿实验 ,研究了饥饿对中华绒螯蟹亲蟹代谢的影响 .结果表明 ,经过 30d的饥饿处理 ,亲蟹的耗氧率降低为摄食状态的 5 0 .0 % ;CO2 排出率减小为摄食状态的 6 3.4 % ;NH3 N排泄率减小为 5 9.1% .耗氧率和CO2 排出率、NH3 N排泄率的变化差别较大 ,耗氧率的减小可分为 4个阶段 ,而CO2 排出率和NH3 N排泄率的减小仅可分 3个阶段 .与此同时 ,亲蟹的标准代谢水平同样受到饥饿的影响 ,饥饿 30d后 ,由 4 .4 5J·g-1·h-1减小为 2 .36J·g-1·h-1;在饥饿亲蟹的代谢中 ,脂肪消耗最多 ,其次是碳水化合物 .  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