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 influence of microheterogeneity on enzyme inactivation kinetics is presented. Examples of different enzymes are given where microheterogeneity has been detected by different techniques. The different statistical models are presented which include the influence of microheterogeneity on enzyme inactivation kinetics and stability. As the microheterogeneity of the enzyme increases, there is a sharper decline in the normalized activity during the initial stages of the deactivation but a greater stability and activity, compared to similar homogeneous enzyme, as the deactivation proceeds. Microheterogeneity makes the deactivation reaction have a higher apparent order of reaction. The implications of microheterogeneity on enzyme inactivations are high lighted by different examples. The analysis provides fresh physical insights into the chemistry, subpopulations, structure, and function of enzymes.  相似文献   

2.
Vanadate-dependent peroxidases contain, in their active center, vanadate covalently attached to histidine in an overall trigonal-bipyramidal array. We describe here the synthesis and characterization of optically active amino alcohols and their vanadium(V) complexes, and we show that the structural models of the active center thus obtained are also functional models for the sulfide-peroxidase activity of the enzyme in heterogeneous catalysis. The heterogeneous systems were obtained by immobilizing the complexes on silica gel and mesoporous silicas, and by aggregation. The following ligands, ligand precursors, and V compounds have been structurally characterized: (R)-(2-phenylethanol)-(R)-1-phenylethylamine (HL(A)), (R,R)-bis[2-phenyl(ethylmethylether)]ammonium chloride ([L(D)]+Cl(-)), the carbasilatranes (R,R)-methoxy{N,N',N'-2,2',3-[bis(1-phenylethanolato)propyl]amino}silane ((R,R)-Si(OMe)L(E)), (R,R)-methoxy-{N,N',N'-1,2',3-[(1-phenylethanolato)-(2-phenylethanolato)propyl]amino}silane ((R,R)-Si(OMe)L(E')), and [VO(L(F))(OSiMe2(t)Bu)], where H2L(F)=ethylbis(2-hydroxy-2-phenylethyl)amine.  相似文献   

3.
The influence of microheterogeneity on enzyme inactivation kinetics is examined. A continuous normal distribution of the thermal activation energy is assumed, and using this, a simple mathematical model is developed to find the activity-time trajectories for a microheterogeneous enzyme. Using an example, the model is used to show the quantitative effects of microheterogeneity such as increased order and stability observed during an enzyme inactivation. Experimental measurement of the extent of microheterogeneity in an enzyme sample is also discussed.  相似文献   

4.
Cellobiase from Aspergillus niger was glycosylated by covalent coupling to cyanogen bromide activated dextran. The conjugated enzyme retained 62% of the original specific activity exhibited by the native cellobiase. The optimum pH as well as the pH stability of the conjugated form remain almost the same as for the native enzyme. Compared to the native enzyme, the conjugated form exhibited a higher optimal reaction temperature and energy of activation, a higher K(m) (Michaelis constant) and lower Vmax (maximal reaction rate), and improved thermal stability. The thermal deactivation of the native and conjugated cellobiase obeyed the first-order kinetics. The calculated half-life values of heat inactivation at 60, 70 and 80 degrees C was 10.7, 6.25, and 4.05 h, respectively, whereas at these temperatures the native enzyme was less stable (half-life of 3.5, 1.69, and 0.83 h, respectively). The deactivation rate constant at 80 degrees C for the conjugated cellobiase is about 7.9 x 10(-2) h-1, which is lower than that of the native enzyme (36.0 x 10(-2) h-1). The activation energy for denaturation of the native enzyme is about 10.58 kcal/mol, which is 7.25 kcal/mol lower than that of the conjugated enzyme. The effect of different surfactants and some metal ions on the activity of the conjugated cellobiase has been investigated.  相似文献   

5.
The inhibition of steroid 5alpha-reductase (5AR) by Delta(1)-4-azasteroids is characterized by a two-step time-dependent kinetic mechanism where inhibitor combines with enzyme in a fast equilibrium, defined by the inhibition constant K(i), to form an initial reversible enzyme-inhibitor complex, which subsequently undergoes a time-dependent chemical rearrangement, defined by the rate constant k(3), leading to the formation of an apparently irreversible, tight-binding enzyme-inhibitor complex (Tian, G., Mook, R. A., Jr., Moss, M. L., and Frye, S. V. (1995) Biochemistry 34, 13453-13459). A detailed kinetic analysis of this process with a series of Delta(1)-4-azasteroids having different C-17 substituents was performed to understand the relationships between the rate of time-dependent inhibition and the affinity of the time-dependent inhibitors for the enzyme. A linear correlation was observed between ln(1/K(i)), which is proportional to the ligand binding energy for the formation of the enzyme-inhibitor complex, and ln(1/(k(3)/K(i))), which is proportional to the activation energy for the inhibition reaction under the second order reaction condition, which leads to the formation of the irreversible, tight-binding enzyme-inhibitor complex. The coefficient of the correlation was -0.88 +/- 0.07 for type 1 5AR and -1.0 +/- 0.2 for type 2 5AR. In comparison, there was no obvious correlation between ln(1/K(i)) and ln(1/k(3)), which is proportional to the activation energy of the second, time-dependent step of the inhibition reaction. These data are consistent with a model where ligand binding energies provided at C-17 of Delta(1)-4-azasteroids is fully expressed to lower the activation energy of k(3)/K(i) with little perturbation of the energy barrier of the second, time-dependent step.  相似文献   

6.
Based on the well-known mechanism describing Michaelis-Menten kinetics, three rate expressions may be developed: the exact solution (Model 1), a rate equation resulting from the pseudo-steady-state assumption (Model 2), and Model 2 with the additional assumption that the amount of free substrate is approximately equal to the total amount of substrate (Model 3). Although Model 1 is the most precise, it must be integrated numerically and it requires three experimentally determined parameters. Models 2 and 3, however, are simpler and require only two parameters. Using dimensionless forms of the three models, we have evaluated the errors in the two simplified models relative to the exact solution using a wide range of parameter values. The choice of model for reactor design depends on the initial substrate to enzyme ratio (alpha(0)), and on the ratio of the Michaelis-Menten constant to the enzyme concentration (sigma). Based on a 2% model error criteria, when alpha(0) > 15 or sigma >/= 100, Model 3 is adequate; if 5 < alpha(0) < 15, or if sigma >/= 10, then Model 2 may be used; and if alpha(0) < 5 and sigma < 10, then the exact solution (Model 1) is required.  相似文献   

7.
Lysosomal alpha-mannosidase is a broad specificity exoglycosidase involved in the ordered degradation of glycoproteins. The bovine enzyme is used as an important model for understanding the inborn lysosomal storage disorder alpha-mannosidosis. This enzyme of about 1,000 amino acids consists of five peptide chains, namely a- to e-peptides and contains eight N-glycosylation sites. The N(497) glycosylation site of the c-peptide chain is evolutionary conserved among LAMANs and is very important for the maintenance of the lysosomal stability of the enzyme. In this work, relying on an approach based on mass spectrometric techniques in combination with exoglycosidase digestions and chemical derivatizations, we will report the detailed structures of the N-glycans and their distribution within six of the eight N-glycosylation sites of the bovine glycoprotein. The analysis of the PNGase F-released glycans from the bovine LAMAN revealed that the major structures fall into three classes, namely high-mannose-type (Fuc(0-1)Glc(0-1)Man(4-9)GlcNAc(2)), hybrid-type (Gal(0-1)Man(4-5)GlcNAc(4)), and complex-type (Fuc(0-1)Gal(0-2)Man(3)GlcNAc(3-5)) N-glycans, with core fucosylation and bisecting GlcNAc. To investigate the exact structure of the N-glycans at each glycosylation site, the peptide chains of the bovine LAMAN were separated using SDS-PAGE and in-gel deglycosylation. These experiments revealed that the N(497) and N(930) sites, from the c- and e-peptides, contain only high-mannose-type glycans Glc(0-1)Man(5-9)GlcNAc(2), including the evolutionary conserved Glc(1)Man(9)GlcNAc(2) glycan, and Fuc(0-1)Man(3-5)GlcNAc(2), respectively. Therefore, to determine the microheterogeneity within the remaining glycosylation sites, the glycoprotein was reduced, carboxymethylated, and digested with trypsin. The tryptic fragments were then subjected to concanavalin A (Con A) affinity chromatography, and the material bound by Con A-Sepharose was purified using reverse-phase high-performance liquid chromatography (HPLC). The tandem mass spectrometry (ESI-MS/MS) and the MALDI analysis of the PNGase F-digested glycopeptides indicated that (1) N(692) and N(766) sites from the d-peptide chain both bear glycans consisting of high-mannose (Fuc(0-1)Man(3-7)GlcNAc(2)), hybrid (Fuc(0-1) Gal(0-1)Man(4-5)GlcNAc(4)), and complex (Fuc(0-1)Gal(0-2)Man(3)GlcNAc(4-5)) structures; and (2) the N(367) site, from the b-peptide chain, is glycosylated only with high-mannose structures (Fuc(0-1)Man(3-5)GlcNAc(2)). Taking into consideration the data obtained from the analysis of either the in-gel-released glycans from the abc- and c-peptides or the tryptic glycopeptide containing the N(367) site, the N(133) site, from the a-peptide, was shown to be glycosylated with truncated and high-mannose-type (Fuc(0-1)Man(4-5)GlcNAc(2)), complex-type (Fuc(0-1)Gal(0-1)Man(3)GlcNAc(5)), and hybrid-type (Fuc(0-1)Gal(0-1)Man(5)GlcNAc(4)) glycans.  相似文献   

8.
The thermal stability of a highly purified preparation of D-amino acid oxidase from Trigonopsis variabilis (TvDAO), which does not show microheterogeneity due to the partial oxidation of Cys-108, was studied based on dependence of temperature (20-60°C) and protein concentration (5-100 µmol L-1). The time courses of loss of enzyme activity in 100 mmol L-1 potassium phosphate buffer, pH 8.0, are well described by a formal kinetic mechanism in which two parallel denaturation processes, partial thermal unfolding and dissociation of the FAD cofactor, combine to yield the overall inactivation rate. Estimates from global fitting of the data revealed that the first-order rate constant of the unfolding reaction (k a) increased 104-fold in response to an increase in temperature from 20 to 60°C. The rate constants of FAD release (k b) and binding (k -b) as well as the irreversible aggregation of the apo-enzyme (k agg) were less sensitive to changes in temperature, their activation energy (E a) being about 52 kJ mol-1 in comparison with an E a value of 185 kJ mol-1 for k a. The rate-determining step of TvDAO inactivation switched from FAD dissociation to unfolding at high temperatures. The model adequately described the effect of protein concentration on inactivation kinetics. Its predictions regarding the extent of FAD release and aggregation during thermal denaturation were confirmed by experiments. TvDAO is shown to contain two highly reactive cysteines per protein subunit whose modification with 5,5'-dithio-bis (2-nitrobenzoic acid) was accompanied by inactivation. Dithiothreitol (1 mmol L-1) enhanced up to 10-fold the recovery of enzyme activity during ion exchange chromatography of technical-grade TvDAO. However, it did not stabilize TvDAO at all temperatures and protein concentrations, suggesting that deactivation of cysteines was not responsible for thermal denaturation.  相似文献   

9.
We study the effect of lipid demixing on the electrostatic interaction of two oppositely-charged membranes in solution, modeled here as an incompressible two-dimensional fluid mixture of neutral and charged mobile lipids. We calculate, within linear and nonlinear Poisson-Boltzmann theory, the membrane separation at which the net electrostatic force between the membranes vanishes, for a variety of different system parameters. According to Parsegian and Gingell, contact between oppositely-charged surfaces in an electrolyte is possible only if the two surfaces have exactly the same charge density (sigma(1) = -sigma(2)). If this condition is not fulfilled, the surfaces can repel each other, even though they are oppositely charged. In our model of a membrane, the lipidic charge distribution on the membrane surface is not homogeneous and frozen, but the lipids are allowed to freely move within the plane of the membrane. We show that lipid demixing allows contact between membranes even if there is a certain charge mismatch, /sigma(1)/ not equal /sigma(2)/, and that in certain limiting cases, contact is always possible, regardless of the value of sigma(1)/sigma(2) (if sigma(1)/sigma(2) < 0). We furthermore find that of the two interacting membranes, only one membrane shows a major rearrangement of lipids, whereas the other remains in exactly the same state it has in isolation and that, at zero-disjoining pressure, the electrostatic mean-field potential between the membranes follows a Gouy-Chapman potential from the more strongly charged membrane up to the point of the other, more weakly charged membrane.  相似文献   

10.
A graphical method is presented for the conformational analysis of the sugar ring in DNA fragments by means of proton-proton couplings. The coupling data required for this analysis consist of sums of couplings, which are referred to as sigma 1' (= J1'2' + J1'2'), sigma 2' (= J1'2' + J2'3' + J2'2'), sigma 2' (= J1'2' + J2'3' + J2'2') and sigma 3' (= J2'3' + J2'3' + J3'4'). These sums of couplings correspond to the distance between the outer peaks of the H1', H2', H2' and H3' [31P] resonances, respectively, (except for sigma 2' and sigma 2' in the case of a small chemical shift difference between the H2' and H2' resonances) and can often be obtained from 1H-NMR spectra via first-order measurement, obviating the necessity of a computer-assisted simulation of the fine structure of these resonances. Two different types of graphs for the interpretation of the coupling data are discussed: the first type of graph serves to probe as to whether or not the sugar ring occurs as a single conformer, and if so to analyze the coupling data in terms of the geometry of this sugar ring. In cases where the sugar ring does not occur as a single conformer, but as a blend of N- and S-type sugar puckers, the second type of graph is used to analyze the coupling data in terms of the geometry and population of the most abundant form. It is shown that the latter type of analysis can be carried out on the basis of experimental values for merely sigma 1',sigma 2' and sigma 2', without any assumptions or restrictions concerning a relation between the geometry of the N- and S-type conformer. In addition, the question is discussed as to how insight can be gained into the conformational purity of the sugar ring from the observed fine structure of the H1' resonance. Finally, a comparison is made between experimental coupling data reported for single-stranded and duplex DNA fragments and covalent RNA-DNA hybrids on the one hand and the predicted couplings and sums of couplings presented in this paper on the other hand.  相似文献   

11.
We link two-allele population models by Haldane and Fisher with Kimura's diffusion approximations of the Wright-Fisher model, by considering continuous-state branching (CB) processes which are either independent (model I) or conditioned to have constant sum (model II). Recent works by the author allow us to further include logistic density-dependence (model III), which is ubiquitous in ecology. In all models, each allele (mutant or resident) is then characterized by a triple demographic trait: intrinsic growth rate r, reproduction variance sigma and competition sensitivity c. Generally, the fixation probability u of the mutant depends on its initial proportion p, the total initial population size z, and the six demographic traits. Under weak selection, we can linearize u in all models thanks to the same master formula u = p + p(1 - p)[g(r)s(r) + g(sigma)s(sigma) + g(c)s(c)] + o(s(r),s(sigma),s(c), where s(r) = r' - r, s(sigma) = sigma-sigma' and s(c) = c - c' are selection coefficients, and g(r), g(sigma), g(c) are invasibility coefficients (' refers to the mutant traits), which are positive and do not depend on p. In particular, increased reproduction variance is always deleterious. We prove that in all three models g(sigma) = 1/sigma and g(r) = z/sigma for small initial population sizes z. In model II, g(r) = z/sigma for all z, and we display invasion isoclines of the 'mean vs variance' type. A slight departure from the isocline is shown to be more beneficial to alleles with low sigma than with high r. In model III, g(c) increases with z like ln(z)/c, and g(r)(z) converges to a finite limit L > K/sigma, where K = r/c is the carrying capacity. For r > 0 the growth invasibility is above z/sigma when z < K, and below z/sigma when z > K, showing that classical models I and II underestimate the fixation probabilities in growing populations, and overestimate them in declining populations.  相似文献   

12.
Plant dispersal, neighbourhood size and isolation by distance   总被引:1,自引:0,他引:1  
Epperson BK 《Molecular ecology》2007,16(18):3854-3865
A theoretical relationship between isolation by distance or spatial genetic structure (SGS) and seed and pollen dispersal is tested using extensive spatial-temporal simulations. Although for animals Wright's neighbourhood size N(e) = 4pisigma(2)(t) has been ascertained also, where sigma(2)(t) is the axial variance of distances between parents and offspring, and it was recently confirmed that N(e) = 4pi(sigma(2)(f) + sigma(2)(m))/2 when dispersal of females and males differ, the situation for plants had not been established. This article shows that for a very wide range of conditions, neighbourhood size defined by Crawford's formula N(e) = 4pi(sigma(2)(s) + sigma(2)(p)/2) fully determines SGS, even when dispersal variances of seed (sigma(2)(s)) and pollen sigma(2)(p)) differ strongly. Further, self-fertilization with rate s acts as zero-distance pollen dispersal, and N(e) = 4pi[sigma(2)(s) + sigma(2)(p)(1 - s)/2] fully determines SGS, for most cases where there are both likely parameter values and substantial SGS. Moreover, for most cases, there is a loglinear relationship, I(1) = 0.587 - 0.117 ln(N(e)), between SGS, as measured by I(1), Moran's coefficient for adjacent individuals, and N(e). However, there are several biologically significant exceptions, namely for very low or large N(e), SGS exceeds the loglinear values. There are also important exceptions to Crawford's formula. First, plants with low seed dispersal, high outcross pollen dispersal and high selfing rate show larger SGS than predicted. Second, in plants with very low (near zero) seed dispersal, selfing decreases SGS, opposite expectations. Finally, in some cases seed dispersal is more critical than pollen dispersal, in a manner inconsistent with Crawford's formula.  相似文献   

13.
This study presents a generic numerical model to simulate the coupled solute and solvent transport in human ovarian tissue sections during addition and removal of chemical additives or cryoprotective agents (CPA). The model accounts for the axial and radial diffusion of the solute (CPA) as well as axial convection of the CPA, and a variable vascular surface area (A) during the transport process. In addition, the model also accounts for the radial movement of the solvent (water) into and out of the vascular spaces. Osmotic responses of various cells within an human ovarian tissue section are predicted by the numerical model with three model parameters: permeability of the tissue cell membrane to water (L(p)), permeability of the tissue cell membrane to the solute or CPA (omega) and the diffusion coefficient of the solute or CPA in the vascular space (D). By fitting the model results with published experimental data on solute/water concentrations within an human ovarian tissue section, I was able to determine the permeability parameters of ovarian tissue cells in the presence of 1.5M solutions of each of the following: dimethyl sulphoxide (DMSO), propylene glycol (PROH), ethylene glycol (EG), and glycerol (GLY), at two temperatures (4 degrees C and 27 degrees C). Modeling Approach 1: Assuming a constant value of solute diffusivity (D = 1.0 x 10(-9) m(2)/sec), the best fit values of L(p) ranged from 0.35 x 10(-14) to 1.43 x 10(-14) m(3)/N-sec while omega ranged from 2.57 x 10(-14) to 70.5 x 10(-14) mol/N-sec. Based on these values of L(p) and omega, the solute reflection coefficient, sigma defined as sigma = 1-omega v(CPA)/L(P) ranged from 0.9961 to 0.9996. Modeling Approach 2: The relative values of omega and sigma from our initial modeling suggest that the embedded ovarian tissue cells are relatively impermeable to all the CPAs investigated (or omega approximately 0 and sigma approximately 1.0). Consequently the model was modified and used to predict the values of L(p) and D assuming omega = 0 and sigma = 1.0. The best fit values of L(p) ranged from 0.44 x 10(-14) to 1.2 x 10(-14) m(3)/N-sec while D ranged from 0.85 x 10(-9) to 2.08 x 10(-9) m(2)/sec. Modeling Approach 3: Finally, the best fit values of D from modeling approach 2 were incorporated into model 1 to re-predict the values of L(p) and omega. It is hoped that the ovarian tissue cell parameters reported here will help to optimize chemical loading and unloading procedures for whole ovarian tissue sections and consequently, tissue cryopreservation procedures.  相似文献   

14.
The dynamic environment within a bioreactor and in the purification equipment is known to affect the activity and yield of enzyme production. The present research focuses on the effect of hydrodynamic flow parameters (average energy dissipation rate, maximum energy dissipation rate, average shear rate, and average normal stress) and the interfacial flow parameters (specific interfacial area and mass transfer coefficient) on the activity of lysozyme. Flow parameters were estimated using CFD simulation based on the k-epsilon approach. Enzyme deactivation was investigated in 0.1, 0.3, 0.57, and 1 m i.d. vessels. Enzyme solution was subjected to hydrodynamic stress using various types of impellers and impeller combinations over a wide range of power consumption (0.03 < P(G)/V < 7, kW/m3). The effects of tank diameter, impeller diameter, blade width, blade angle, and the number of blades on the extent of deactivation were investigated. At equal value of P(G)/V, epsilon(max), and gamma(avg), the extent of deactivation was dramatically different for different impeller types. The extent of deactivation was found to correlate well with the average turbulent normal stress and the mass transfer coefficient.  相似文献   

15.
The venom of Naja nigricolis was found to contain a high level of the enzyme aryl sulfatase. The enzyme was isolated from the venom of N. nigriclois and purified to electrophoretic homogeneity by gel chromatography on Sephadex G-100, DEAE-cellulose, and phenyl-sepharose columns. The enzyme was optimally active at pH 5 and 40 degrees C. Arrhenius plot for the determination of the activation energy (E(a)) gave the value 25 kJ/mol with a half-life (t(1/2)) of 5 min at 50 degrees C. It was highly activated by Fe(2+) and Ca(2+) and inhibited by Co(2+) and Mn(2+). The enzyme catalyzed the hydrolysis of the fluorescent compound methylumbelliferyl-sulfate (MU-SO(4)). Double reciprocal plots of initial velocity data, using MU-SO(4) as substrate, gave a K(M) value of 110 microM and V(max) of 225 micromol min(-1) x mg(-1). N. nigricolis Aryl sulphatase also hydrolyzed chondroitin-4-sulphate. It was inhibited competitively by N-acetyl glucosamine sulfate (GlcNAc-SO(4)), glucose-6-sulfate (Glc-6-SO(4)), and glucose 1-sulfate (Glc-1-SO(4)). Extrapolated inhibition binding constants (K(i)) gave the values of 3, 25, and 315 microM for GlcNAc-SO(4), Glc-6-SO(4), and Glc-1-SO(4) respectively.  相似文献   

16.
Studies on the interaction of calcineurin with its activator, calmodulin, showed that the 1:1 complex is the activated species. Concomitant with activation, a time-dependent deactivation of the phosphatase was observed. The process followed first order kinetics and was dependent on the presence of both Ca2+ and calmodulin. The deactivation rate constant at pH 7.6 and 30 degrees C was 0.06 min-1, which was increased by the substrate, p-nitrophenylphosphate (Km = 11 mM), to 0.47 min-1. PPi and nucleotides inhibited the enzyme competitively and accelerated the deactivation. The first order rate constant was increased to 2.3 min-1 by PPi (Ki = 55 microM) and to 8.0 min-1 by ADP (Ki = 0.94 mM). A theory dealing with the deactivation (applicable to chemical modification, etc.) of an enzyme in the absence and presence of various ligands is presented. The deactivated enzyme remained bound to calmodulin and was not reactivated by dissociation-reassociation of the calcineurin-calmodulin complex. Calcineurin was found to contain covalently bound phosphate; however, no difference in its content was detected upon deactivation, indicating that self-dephosphorylation was not involved. The deactivation could be reversed, as well as prevented, by divalent metal ions such as Ni2+ and Mn2+. Atomic absorption spectroscopy revealed nearly stoichiometric amounts of tightly bound Fe and Zn (but little other ions) on purified calcineurin, which remained bound during the calmodulin-dependent deactivation; removal of tightly bound metals is, therefore, not the cause of deactivation. Our results indicate that calcineurin is a metallophosphatase and not simply a Ca2+- and calmodulin-stimulated enzyme. In addition to the nondissociable Zn and Fe and the Ca2+ bound to the B subunit and calmodulin, the enzyme requires a divalent metal ion for structural stability and full activity.  相似文献   

17.
In standard models of quantitative traits, genotypes are assumed to differ in mean but not variance of the trait. Here we consider directional selection for a quantitative trait for which genotypes also confer differences in variability, viewed either as differences in residual phenotypic variance when individual loci are concerned or as differences in environmental variability when the whole genome is considered. At an individual locus with additive effects, the selective value of the increasing allele is given by ia/sigma + 1/2 ixb/sigma2, where i is the selection intensity, x is the standardized truncation point, sigma2 is the phenotypic variance, and a/sigma and b/sigma2 are the standardized differences in mean and variance respectively between genotypes at the locus. Assuming additive effects on mean and variance across loci, the response to selection on phenotype in mean is isigma2(Am)/sigma + 1/2 ixcov(Amv)/sigma2 and in variance is icov(Amv)/sigma + 1/2 ixsigma2(Av)/sigma2, where sigma2(Am) is the (usual) additive genetic variance of effects of genes on the mean, sigma2(Av) is the corresponding additive genetic variance of their effects on the variance, and cov(Amv) is the additive genetic covariance of their effects. Changes in variance also have to be corrected for any changes due to gene frequency change and for the Bulmer effect, and relevant formulae are given. It is shown that effects on variance are likely to be greatest when selection is intense and when selection is on individual phenotype or within family deviation rather than on family mean performance. The evidence for and implications of such variability in variance are discussed.  相似文献   

18.
The redox behaviour of the Ni(III)/Ni(II) transition in hydrogenase from Chromatium vinosum is described and compared with the redox behaviour of the nickel ion in the F420-nonreducing hydrogenase from Methanobacterium thermoautotrophicum. Analogous to the situation in the oxidised hydrogenase of Desulfovibrio gigas (Fernandez, V.M., Hatchikian, E.C., Patil, D.S. and Cammack, R. (1986) Biochim. Biophys. Acta 883, 145-154), the C. vinosum enzyme can also exist in two forms: the 'unready' form (EPR characteristics of Ni(III): gx,y,z = 2.32, 2.24, 2.01) and the 'ready' form (EPR characteristics Ni(III): gx,y,z = 2.34, 2.16, 2.01). Like in the oxidised enzyme of M. thermoautotrophicum the Ni(III)/Ni(II) transition for the unready form titrated completely reversible (both at pH 6.0 and pH 8.0). In contrast, the reversibility of the Ni(III)/Ni(II) transition in the ready enzyme was strongly dependent on pH and temperature. At pH 6.0 and 2 degrees C reduction of Ni(III) in ready enzyme was completely irreversible, whereas at pH 8.0 and 30 degrees C Ni(III) in both ready and unready enzyme titrated with E0' = -115 mV (n = 1). Hampered redox equilibration between the ready enzyme and the mediating dyes is interpreted in terms of an obstruction of the electron transfer from nickel at the active site to the artificial electron acceptors in solution. The origin of this obstruction might be related to possible changes in the protein structure induced by the activation process. The E0'-value of the Ni(III)/Ni(II) equilibrium was pH sensitive (-60 mV/delta pH) indicating that reduction of nickel is coupled to a protonation. A similar pH-dependence was observed for the titration of the spin-spin interaction of Ni(III) and a special form of the [3Fe-4S]+ cluster (E0' = +150 mV, pH 8.0, 30 degrees C). Redox equilibration of this coupling was extremely sensitive to pH and temperature. The uncoupled [3Fe-4S]+ cluster titrated pH-independently with E0' = -10 mV (pH 8.0, 30 degrees C).  相似文献   

19.
Plesiomonas shigelloides is a Gram-negative bacterium associated with waterborne infections, which is common in tropical and subtropical habitats. Contrary to the unified antigenic classification of P. shigelloides, data concerning the structure and activity of their lipopolysaccharides (LPS and endotoxin) are limited. This study completes the structural investigation of phenol- and water-soluble fractions of P. shigelloides O74 (strain CNCTC 144/92) LPS with the emphasis on lipid A heterogeneity, describing the entire molecule and some of its biological in vitro activities. Structures of the lipid A and the affinity-purified decasaccharide obtained by de-N,O-acylation of P. shigelloides O74 LPS were elucidated by chemical analysis combined with electrospray ionization multiple-stage mass spectrometry (ESI-MS(n)), MALDI-TOF MS, and NMR spectroscopy. Lipid A of P. shigelloides O74 is heterogeneous, and three major forms have been identified. They all were asymmetric, phosphorylated, and hexaacylated, showing different acylation patterns. The beta-GlcpN4P-(1-->6)-alpha-GlcpN1P disaccharide was substituted with the primary fatty acids: (R)-3-hydroxytetradecanoic acid [14:0(3-OH)] at N-2 and N-2' and (R)-3-hydroxydodecanoic acid [12:0(3-OH)] at O-3 and O-3'. The heterogeneity among the three forms (I-III) of P. shigelloides O74 lipid A was attributed to the substitution of the acyl residues at N-2' and O-3' with the secondary acyls: (I) cis-9-hexadecenoic acid (9c-16:1) at N-2' and 12:0 at O-3', (II) 14:0 at N-2' and 12:0 at O-3', and (III) 12:0 at N-2' and 12:0 at O-3'. The pro-inflammatory cytokine-inducing activities of P. shigelloides O74 LPS were similar to those of Escherichia coli O55 LPS.  相似文献   

20.
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