首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Thermodynamics of phospholipid-sucrose interactions.   总被引:2,自引:1,他引:1       下载免费PDF全文
The effect of 0-1.0 M sucrose on the phase-transition properties of 1,2-dipalmitoyl-3-sn-phosphatidylcholine (1,2-DPPC) was examined by high-sensitivity differential scanning calorimetry at a scan rate of 0.1 K min-1. Increasing the concentration of sucrose caused a small, but experimentally significant, increase in the temperature (Tm) of maximal excess apparent specific heat (Cmax) and in delta T 1/2 (the transition width at 1/2 Cmax), a reduction in Cmax, and a small decrease (approximately 8-10% at 1.0 M sucrose compared with 0 M sucrose) in the calorimetric enthalpy (delta Hcal) of the gel-to-liquid crystalline transition. The calorimetric parameters of the pretransition of 1,2-DPPC were not significantly affected by sucrose in the concentration range examined, except there was a 1.0 degree C increase in the temperature (Tp) of maximal excess apparent specific heat in the presence of 1.0 M sucrose. The results are discussed in terms of the possible molecular mechanisms that could have caused the observed changes and are contrasted with the results obtained by C. -H. Chen et al. (1981, Biophys. J., 36:359-367).  相似文献   

2.
Oxidized rubredoxin from Clostridium pasteurianum has been investigated by magnetic circular dichroism (MCD) spectroscopy over the temperature range 1.5 to 150 K and at magnetic fields between 0 and 4.5 tesla. The results show that studies of the temperature and field dependence of MCD transitions afford insight into the polarization of electronic transitions for ground states with large g-value anisotropy, in addition to estimates of ground-state g values and zero-field splitting parameters. In agreement with the assignment made by Eaton and Lovenberg (Eaton, W.A. and Lovenberg, W. (1973) in Iron-Sulfur Proteins, Vol. II (Lovenberg, W., ed.), pp. 131-162, Academic Press, New York), the ultraviolet-visible spectrum of oxidized rubredoxin is assigned to two S----Fe(III) charge transfer transitions (both 6A1----6T2 under tetrahedral symmetry), each spanning a range of 650-430 nm and 430-330 nm, respectively. The observed splitting in each of these transitions is attributed to a predominant axial distortion in the excited state resulting in effective D2d symmetry.  相似文献   

3.
The properties of folinate and 5-methyltetrahydrofolate (5-CH(3)-H(4)PteGlu) transport mechanism of Pediococcus cerevisiae were studied. The uptake was dependent on temperature, pH (optimum for both compounds at pH 6.0), and glucose. Iodoacetate, potassium fluoride, and sodium azide inhibited the uptake. 5-CH(3)-H(4)-PteGlu was apparently not metabolized but folinate was metabolized. Metabolism of folinate was reduced by preincubation of cells with fluorodeoxyuridine. The transport system for folinate and 5-CH(3)-H(4)PteGlu were specific for the l-isomers. Pteroylglutamate, aminopterin, and amethopterin did not interfere with the uptake. Tetrahydrofolate competed with the uptake of folinate. The transport of folinate and 5-CH(3)-H(4)PteGlu at 37 C conformed to Michaelis-Menten kinetics; apparent K(m) for both compounds was 4.0 x 10(-7)m, and the V(max) for folinate was 1.0 x 10(-10) moles per min per mg (dry weight) and for 5-CH(3)-H(4)PteGlu it was 1.6 x 10(-10) moles per min per mg (dry weight). Both compounds accumulated in the intracellular pool at a concentration about 80- to 140-fold higher than that in the external medium. Folinate inhibited competitively the uptake of 5-CH(3)-H(4)PteGlu with a K(i) of 0.4 x 10(-7)m. Unlike 5-CH(3)-H(4)PteGlu, which accumulated only at 37 C, folinate was also taken up at 0 C by a glucose- and temperature-independent process, which was not affected by the metabolic inhibitors mentioned above. Since at 0 C the intracellular concentration of folinate was also considerably higher than the external, binding of the substrate to some cellular component is assumed. The finding of an efficient transport system for l-5-CH(3)-H(4)PteGlu is of special interest, since this compound has no growth-promoting activity for P. cerevisiae.  相似文献   

4.
The thermodynamics of the conversion of aqueous xylose to xylulose has been investigated using high-pressure liquid chromatography (HPLC) and microcalorimetry. The reaction was carried out in aqueous phosphate buffer over the pH range 6.8-7.4 using solubilized glucose isomerase with MgSO(4) as a cofactor. The temperature range over which this reaction was investigated was 298.15-342.15 K. A combined analysis of both the HPLC and microcalorimetric data leads to the following results at 298.15 K for the conversion process: DeltaG degrees = 4389 +/- 31 J mol(-1), DeltaH degrees = 16090 +/- 670 J mol(-1), and DeltaC(p) degrees = 40 +/- 23 J mol(-1) K(-1). The temperature dependence of the equilibrium constant for the reaction is expressed as R ln K = -4389/298.15 +16090[(1/298.15)-(1/T)]+40[(298.15/T)-1 + ln(T/298.15)]. Comparisons are made with literature data.  相似文献   

5.
Data are reported for T(m), the temperature midpoint of the thermal unfolding curve, of ribonuclease A, versus pH (range 2-9) and salt concentration (range 0-1 M) for two salts, Na(2)SO(4) and NaCl. The results show stabilization by sulfate via anion-specific binding in the concentration range 0-0.1 M and via the Hofmeister effect in the concentration range 0.1-1.0 M. The increase in T(m) caused by anion binding at 0.1 M sulfate is 20 degrees at pH 2 but only 1 degree at pH 9, where the net proton charge on the protein is near 0. The 10 degrees increase in T(m) between 0.1 and 1.0 M Na(2)SO(4), caused by the Hofmeister effect, is independent of pH. A striking property of the NaCl results is the absence of any significant stabilization by 0.1 M NaCl, which indicates that any Debye screening is small. pH-dependent stabilization is produced by 1 M NaCl: the increase in T(m) between 0 and 1.0 M is 14 degrees at pH 2 but only 1 degree at pH 9. The 14 degree increase at pH 2 may result from anion binding or from both binding and Debye screening. Taken together, the results for Na(2)SO(4) and NaCl show that native ribonuclease A is stabilized at low pH in the same manner as molten globule forms of cytochrome c and apomyoglobin, which are stabilized at low pH by low concentrations of sulfate but only by high concentrations of chloride.  相似文献   

6.
The sulfur-containing amino acid, methionine, has a role in the physiological environment because of its strong interactions with metals. To understand these interactions of metals with methionine, one needs reliable dissociation constants for the protonated methionine species (NH(3)(+)CH(CH(2)CH(2)SCH(3))COOH; H(2)B(+)). The values of stoichiometric dissociation constants, pK(i)*, for protonated methionine species (H(2)B(+) if H(+)+HB, K(1); HB if H(+)+B(-), K(2)) were determined from potentiometric measurements in NaCl solutions as a function of ionic strength, 0.25-6.0 mol (kg H(2)O)(-1) and temperature (5-45 degrees C). The results were extrapolated to pure water using the Pitzer equations to estimate the activity of H(+), H(2)B(+), HB and B(-) as a function of ionic strength and temperature. The resulting thermodynamic values of K(1) and K(2) were fit to the equations (T/K): ln K(1)=69.0013-3496.58/(T/K)-10.9153 ln (T/K); ln K(2)=116.4162-10638.02/(T/K)-18.0553 ln (T/K) with standard errors of 0.003 and 0.033, respectively, for ln K(1)* and ln K(2)*. Pitzer interaction parameters (lambda(HB-Na) and zeta(HB-Na-Cl)) for the neutral HB were determined from literature data. The Pitzer parameters (beta(0)(H(2)BCl), beta(1)(H(2)BCl) and C(phi)(H(2)BCl)) for the interactions of H(2)B(+) with Cl(-) and Na(+) with and B(2-) (beta(0)(NaB), beta(1)(NaB) and C(phi)(NaB)) were also determined. These coefficients can be used to make reasonable estimates of the activity coefficients of methionine species and the pK(i)(*) for the dissociation of methionine in physiological solutions, composed mostly of NaCl over a wide range of temperature and ionic strength.  相似文献   

7.
Molecular dynamics simulations (NPT ensembles, 1 atm) using the all atom force field AMB99C (F. A. Momany and J. L. Willett, Carbohydrate Research, Vol. 326, pp 194-209 and 210-226), are applied to a periodic cell containing ten maltodecaose fragments and TIP3P water molecules. Simulations were carried out at 25 K intervals over a range of temperatures above and below the expected glass transition temperature, T(g), for different water concentrations. The amorphous cell was constructed through successive dynamic equilibration steps at temperatures above T(g) and the temperature lowered until several points of reduced slope (1/T vs volume) were obtained. This procedure was carried out at each hydration level. Each dynamics simulation was continued until the volume remained constant without up or down drift for at least the last 100 ps. For a given temperature, most simulations required 400-600 ps to reach an equilibrium state, but longer times were necessary as the amount of water in the cell was reduced. A total of more than 30 ns of simulations were required for the complete study. The T(g) for each hydrated cell was taken as that point at which a discontinuity in slope of the volume (V), potential energy (PE), or density (rho) vs 1/T was observed. The average calculated T(g) values were 311, 337, 386, and 477 K for hydration levels of 15.8, 10, 5, and 1%, respectively, in generally good agreement with experimental values. The T(g) for anhydrous amylose is above the decomposition temperature for carbohydrates and so cannot be easily measured. However, it has also been difficult to obtain a value of T(g) for anhydrous amylose using simulation methods. Other molecular parameters such as end-to-end distances, mean square distributions, and pair distributions are discussed.  相似文献   

8.
The thermodynamics of the equilibria between aqueous ribose, ribulose, and arabinose were investigated using high-pressure liquid chromatography and microcalorimetry. The reactions were carried out in aqueous phosphate buffer over the pH range 6.8-7.4 and over the temperature range 313.15-343.75 K using solubilized glucose isomerase with either Mg(NO3)2 or MgSO4 as cofactors. The equilibrium constants (K) and the standard state Gibbs energy (delta G degrees) and enthalpy (delta H degrees) changes at 298.15 K for the three equilibria investigated were found to be: ribose(aq) = ribulose(aq) K = 0.317, delta G degrees = 2.85 +/- 0.14 kJ mol-1, delta H degrees = 11.0 +/- 1.5 kJ mol-1; ribose(aq) = arabinose(aq) K = 4.00, delta G degrees = -3.44 +/- 0.30 kJ mol-1, delta H degrees = -9.8 +/- 3.0 kJ mol-1; ribulose(aq) = arabinose(aq) K = 12.6, delta G degrees = -6.29 +/- 0.34 kJ mol-1, delta H degrees = -20.75 +/- 3.4 kJ mol-1. Information on rates of the above reactions was also obtained. The temperature dependencies of the equilibrium constants are conveniently expressed as R in K = -delta G degrees 298.15/298.15 + delta H degrees 298.15[(1/298.15)-(1/T)] where R is the gas constant (8.31441 J mol-1 K-1) and T the thermodynamic temperature.  相似文献   

9.
We studied the energy landscape of the peptide Ace-GEWTYDDATKTFTVTE-Nme, taken from the C-terminal fragment (41-56) of protein G, in explicit aqueous solution by a highly parallel replica-exchange approach that combines molecular dynamics trajectories with a temperature exchange Monte Carlo process. The combined trajectories in T and configurational space allow a replica to overcome a free energy barrier present at one temperature by increasing T, changing configurations, and cooling in a self-regulated manner, thus allowing sampling of broad regions of configurational space in short (nanoseconds) time scales. The free energy landscape of this system over a wide range of temperatures shows that the system preferentially adopts a beta hairpin structure. However, the peptide also samples other stable ensembles where the peptide adopts helices and helix-turn-helix states, among others. The helical states become increasingly stable at low temperatures, but are slightly less stable than the beta turn ensemble. The energy landscape is rugged at low T, where substates are separated by large energy barriers. These barriers disappear at higher T (approximately 330 K), where the system preferentially adopts a "molten globule" state with structures similar to the beta hairpin.  相似文献   

10.
Proteins with a BLUF (sensor of blue light using flavin adenine dinucleotide) domain represent a newly recognized class of photoreceptors that is widely distributed in the genomes of photosynthetic bacteria, cyanobacteria, and Euglena. Recently, Okajima et al. [Okajima, K., Yoshihara, S., Geng, X., Katayama, M. and Ikeuchi, M. (2003) Plant Cell Physiol. 44 (Suppl), 162] purified BLUF protein Tll0078 encoded in the genome of thermophilic cyanobacterium Thermosynechococcus elongatus BP-1 by expressing the protein in Escherichia coli. We investigated the photocycle of Tll0078 by measuring the picosecond fluorescence kinetics, transient absorption changes, and the UV-visible absorption spectra at 10 to 330 K. The absorption spectrum of the FAD moiety of Tll0078 showed a 10-nm red shift upon illumination at 278-330 K. The quantum efficiency of the formation of the red-shifted form was 29%. Illumination at 10 K, on the other hand, caused only a 5-nm red shift in about one-half of the protein population. The 5-nm-shifted form was stable at 10 K. The 5-nm red-shifted form was converted into the 10-nm red-shifted form at 50-240 K upon warming in the dark. At room temperature, the 10-nm red-shifted final product appeared within 10 ns after laser flash excitation. The lifetime of the fluorescence of FAD was found to be 120 ps at room temperature. These results reveal a fast and efficient photoconversion process from the singlet-excited state to the final product at room temperature. A photocycle of BLUF protein is proposed that includes the 5-nm red-shifted intermediate form as the precursor for the 10-nm red-shifted final product. The temperature dependence of each step of the photocycle is also discussed.  相似文献   

11.
The effect of simulated altitude erythrocythemia on hemoglobin flow rate and maximal O2 uptake (VO2max) was determined for nine women sea-level residents. Test conditions included normoxia and normobaric hypoxia (16% O2-84% N2). Cycle tests were performed under normoxia (T1-N) and hypoxia (T1-H) at prereinfusion control and under hypoxia 48 h after a placebo infusion (T2-H) and 48 h after autologous infusion of 334 ml of erythrocytes (T3-H). Hematocrit (38.1-44.9%) and hemoglobin concentration (12.7-14.7 g.dl-1) increased from control to postreinfusion. At peak exercise, VO2max decreased from T1-N (2.40 l.min-1) to T1-H (2.15 l.min-1) then increased at T3-H (2.37 l.min-1). Maximal arterial-mixed venous O2 difference decreased from T1-N to T1-H and increased at T3-H. Cardiac output (Q), stroke volume, heart rate, and total peripheral resistance during maximal exercise were unchanged from T1-N through T3-H. Hemoglobin flow rate (Hb flow) at maximum did not change from T1-N to T1-H but increased at T3-H. When compared with submaximal values for T1-N, VO2 was unchanged at T1-H and T3-H; Q increased at T1-H and decreased at T3-H; arterial-mixed venous O2 difference decreased at T1-H and increased at T3-H; Hb flow did not change at T1-N but increased at T3-H. For young women, simulated altitude erythrocythemia increased peak Hb flow and decreased physiological altitude (227.8 m) but did not affect maximum cardiac output (Qmax).  相似文献   

12.
The thermodynamics of the enzymatic hydrolysis of cellobiose, gentiobiose, isomaltose, and maltose have been studied using both high pressure liquid chromatography and microcalorimetry. The hydrolysis reactions were carried out in aqueous sodium acetate buffer at a pH of 5.65 and over the temperature range of 286 to 316 K using the enzymes beta-glucosidase, isomaltase, and maltase. The thermodynamic parameters obtained for the hydrolysis reactions, disaccharide(aq) + H2O(liq) = 2 glucose(aq), at 298.15 K are: K greater than or equal to 155, delta G0 less than or equal to -12.5 kJ mol-1, and delta H0 = -2.43 +/- 0.31 kJ mol-1 for cellobiose; K = 17.9 +/- 0.7, delta G0 = -7.15 +/- 0.10 kJ mol-1 and delta H0 = 2.26 +/- 0.48 kJ mol-1 for gentiobiose; K = 17.25 +/- 0.7, delta G0 = -7.06 +/- 0.10 kJ mol-1, and delta H0 = 5.86 +/- 0.54 kJ mol-1 for isomaltose; and K greater than or equal to 513, delta G0 less than or equal to -15.5 kJ mol-1, and delta H0 = -4.02 +/- 0.15 kJ mol-1 for maltose. The standard state is the hypothetical ideal solution of unit molality. Due to enzymatic inhibition by glucose, it was not possible to obtain reliable values for the equilibrium constants for the hydrolysis of either cellobiose or maltose. The entropy changes for the hydrolysis reactions are in the range 32 to 43 J mol-1 K-1; the heat capacity changes are approximately equal to zero J mol-1 K-1. Additional pathways for calculating thermodynamic parameters for these hydrolysis reactions are discussed.  相似文献   

13.
Carboxylated poly(styrene/acrylamide) (P(St/AAm)-H) microspheres with different acrylamide contents were prepared by emulsifier-free emulsion polymerization. Thermus thermophilus holo-chaperonin (cpn) was covalently immobilized onto these microspheres with high yield. The T. thermophilus holo-cpn-immobilized microspheres were used for refolding of guanidine hydrochloride (Gdn-HCl)-denatured enzymes and showed sufficiently high ability to facilitate refolding of Leuconostoc mesenteroides glucose-6-phosphate dehydrogenase (G6PD) and pig heart lactate dehydrogenase (LDH) at 30 degrees C and Bacillus stearothermophilus LDH at 60 degrees C. The specific ability of T. thermophilus holo-cpn-immobilized microspheres increased with increasing immobilization amount and reached plateau at around 10-15 mg/m(2). When the immobilization amounts of T. thermophilus holo-cpn were approximately 10 mg/m(2), the microspheres retained sufficiently high ability to facilitate protein refolding during repeated use. Therefore, the P(St/AAm)-H microspheres on which approximately 10 mg/m(2) of T. thermophilus holo-cpn is immobilized are very effective for refolding of various (Gdn-HCl)-denatured enzymes over a wide temperature range.  相似文献   

14.
Specific recognition of the mRNA 5' cap by eukaryotic initiation factor eIF4E is a rate-limiting step in the translation initiation. Fluorescence spectroscopy and high-sensitivity isothermal titration calorimetry were used to examine the thermodynamics of eIF4E binding to a cap-analogue, 7-methylGpppG. A van't Hoff plot revealed nonlinearity characterized by an unexpected, large positive molar heat capacity change (DeltaC(degree)(p) = +1.92 +/- 0.93 kJ.mol(-1).K(-1)), which was confirmed by direct ITC measurements (DeltaC(degree)(p) = +1.941 +/- 0.059 kJ.mol(-1).K(-1)). This unique result appears to come from an extensive additional hydration upon binding and charge-related interactions within the binding site. As a consequence of the positive DeltaC(degree)(p), the nature of the thermodynamic driving force changes with increasing temperature, from enthalpy-driven and entropy-opposed, through enthalpy- and entropy-driven in the range of biological temperatures, into entropy-driven and enthalpy-opposed. Comparison of the van't Hoff and calorimetric enthalpy values provided proof for the ligand protonation at N(1) upon binding, which is required for tight stabilization of the cap-eIF4E complex. Intramolecular self-stacking of the dinucleotide cap-analogue was analyzed to reveal the influence of this coupled process on the thermodynamic parameters of the eIF4E-mRNA 5' cap interaction. The temperature-dependent change in the conformation of 7-methylGpppG shifts significantly the intrinsic DeltaH(degree)(0) = -72.9 +/- 4.2 kJ.mol(-1) and DeltaS(degree)(0) = -116 +/- 58 J.mol(-1).K(-1) of binding to the less negative resultant values, by DeltaH(degree)(sst) = +9.76 +/- 1.15 kJ.mol(-1) and DeltaS(degree)(sst) = +24.8 +/- 2.1 J.mol(-1).K(-1) (at 293 K), while the corresponding DeltaC(degree)(p)(sst) = -0.0743 +/- 0.0083 kJ.mol(-1).K(-1) is negligible in comparison with the total DeltaC(degree)(p) .  相似文献   

15.
Glycerol has been shown to lower the heat denaturation temperature (T(m)) of dehydrated lysozyme while elevating the T(m) of hydrated lysozyme (. J. Pharm. Sci. 84:707-712). Here, we report an in situ elastic neutron scattering study of the effect of glycerol and hydration on the internal dynamics of lysozyme powder. Anharmonic motions associated with structural relaxation processes were not detected for dehydrated lysozyme in the temperature range of 40 to 450K. Dehydrated lysozyme was found to have the highest T(m) by. Upon the addition of glycerol or water, anharmonicity was recovered above a dynamic transition temperature (T(d)), which may contribute to the reduction of T(m) values for dehydrated lysozyme in the presence of glycerol. The greatest degree of anharmonicity, as well as the lowest T(d), was observed for lysozyme solvated with water. Hydrated lysozyme was also found to have the lowest T(m) by. In the regime above T(d), larger amounts of glycerol lead to a higher rate of change in anharmonic motions as a function of temperature, rendering the material more heat labile. Below T(d), where harmonic motions dominate, the addition of glycerol resulted in a lower amplitude of motions, correlating with a stabilizing effect of glycerol on the protein.  相似文献   

16.
The complex [Fe2S2(S2-o-xylyl)2]2- in DMF (dimethylformamide), DMSO (dimethylsulphoxide) or a 1:1 DMF/DMSO mixture, a model for the chromophore in the 2Fe-2S proteins (ferredoxins), has been reduced and studied by conventional EPR over a temperature range. The low-field feature of the spectrum, Hz, has been computer simulated in order to analyse the lineshape in terms of a convolution product of Lorentzian and Gaussian distributions. The Gaussian contribution to the linewidth and a fixed part of the Lorentzian contribution, which is a function of the solvent and the way it freezes, were measured at a low temperature (less than or equal to 100 K) and subtracted from the linewidths in the higher-temperature range (130-200 K). The variable Lorentzian contribution thus obtained was related to spin-lattice relaxation times. The spin-lattice relaxation times of the sample having 1:1 DMSO/DMF solvent were measured in the range 6 to 11 K by the saturating pulse technique and in the range 10 to 65 K by continuous saturation methods. Up to 65 K the results follow the law 1/T1 alpha T4.5, a relationship which is not readily interpreted in terms of a simple Debye model. At higher temperatures the results may be interpreted in terms either of a dominant Orbach mechanism involving excited states at approx. 900 +/- 50 cm-1 (DMSO, DMF) or 770 +/- 50 cm-1 (1:1 DMSO/DMF), or of a Raman process in which 1/T1 alpha T7.5. The former is compatible with the two-phonon process already described in some ferredoxins, especially those with little anisotropy (gy - gx approximately 0.0) which have characteristically high [J] values.  相似文献   

17.
赵晨宇  李新畅  崔娟  高宇  史树森 《昆虫学报》1950,63(9):1108-1116
【目的】明确自然变温环境对甘蓝夜蛾Mamestra brassicae生长发育和繁殖的影响,深入研究其对环境温度的适应性。【方法】在13~25℃(日平均19℃), 16~28℃(日平均22℃), 19~31℃(日平均25℃), 22~34℃(日平均28℃)和25~37℃(日平均31℃) 5个梯度变温条件下,以大豆Glycine max植株叶片为寄主材料饲养甘蓝夜蛾卵,测定其各虫态发育历期、发育速率、成虫繁殖力及发育起点温度和有效积温。【结果】变温范围为13~25℃时甘蓝夜蛾发育历期最长,世代发育历期为65.93 d,显著长于其他变温处理。且随温度升高,其发育历期缩短,变温范围为22~34℃时,该虫发育历期最短,世代发育历期为38.46 d,显著短于其他变温处理。在变温范围为25~37℃时,该虫不能正常完成个体发育。在日平均温度(T)19~28℃范围内(最大温差12℃),甘蓝夜蛾卵、幼虫及蛹期的发育速率随温度升高而加快,且各个虫态发育速率(V)拟合方程均符合线性方程模型:V卵期=0.125+0.048T, V幼虫期=0.023+0.012T, V蛹期0.027+0.013T, V成虫=0.073+0.47T。甘蓝夜蛾雌雄成虫的寿命随着日平均温度的升高而逐渐缩短,雌雄成虫寿命在日变温范围13~25℃时最长,分别为7.91 d和8.00 d;在变温范围22~34℃时最短,分别为3.00 d和3.57 d。甘蓝夜蛾卵、幼虫、蛹、成虫发育起点温度分别为7.98, 6.54, 9.36和10.78℃,有效积温依次为87.00, 607.36, 351.51和108.52 d·℃。16~28℃的变温范围更适合甘蓝夜蛾种群的生存与繁殖,其种群趋势指数I为117.81。【结论】甘蓝夜蛾属于偏低温适应性害虫,对高温环境适应能力较低。研究结果为进一步研究甘蓝夜蛾自然种群发生规律及其发生期、发生量预测预报提供了科学依据。  相似文献   

18.
Agrawal N  Hong B  Mihai C  Kohen A 《Biochemistry》2004,43(7):1998-2006
The enzyme thymidylate synthase (TS) catalyzes a complex reaction that involves forming and breaking at least six covalent bonds. The physical nature of the hydride transfer step in this complex reaction cascade has been studied by means of isotope effects and their temperature dependence. Competitive kinetic isotope effects (KIEs) on the second-order rate constant (V/K) were measured over a temperature range of 5-45 degrees C. The observed H/T ((T)V/K(H)) and D/T ((T)V/K(D)) KIEs were used to calculate the intrinsic KIEs throughout the temperature range. The Swain-Schaad relationships between the H/T and D/T V/K KIEs revealed that the hydride transfer step is the rate-determining step at the physiological temperature of Escherichia coli (20-30 degrees C) but is only partly rate-determining at elevated and reduced temperatures. H/D KIE on the first-order rate constant k(cat) ((D)k = 3.72) has been previously reported [Spencer et al. (1997) Biochemistry 36, 4212-4222]. Additionally, the Swain-Schaad relationships between that (D)k and the V/K KIEs reported here suggested that at 20 degrees C the hydride transfer step is the rate-determining step for both rate constants. Intrinsic KIEs were calculated here and were found to be virtually temperature independent (DeltaE(a) = 0 within experimental error). The isotope effects on the preexponential Arrhenius factors for the intrinsic KIEs were A(H)/A(T) = 6.8 +/- 2.8 and A(D)/A(T) = 1.9 +/- 0.25. Both effects are significantly above the semiclassical (no-tunneling) predicted values and indicate a contribution of quantum mechanical tunneling to this hydride transfer reaction. Tunneling correction to transition state theory would predict that these isotope effects on activation parameters result from no energy of activation for all isotopes. Yet, initial velocity measurements over the same temperature range indicate cofactor inhibition and result in significant activation energy on k(cat) (4.0 +/- 0.1 kcal/mol). Taken together, the temperature-independent KIEs, the large isotope effects on the preexponential Arrhenius factors, and a significant energy of activation all suggest vibrationally enhanced hydride tunneling in the TS-catalyzed reaction.  相似文献   

19.
Inhibition of bovine brain hexokinase by its product, glucose 6-phosphate, is considered to be a major regulatory step in controlling the glycolytic flux in the brain. Investigations on the molecular basis of this regulation, i.e. allosteric or product inhibition, have led to various proposals. Here, we attempt to resolve this issue by ascertaining the location of the binding sites for glucose and glucose 6-phosphate on the enzyme with respect to a divalent-cation-binding site characterized previously [Jarori, G. K., Kasturi, S. R. & Kenkare, U. W. (1981) Arch. Biochem. Biophys. 211, 258-268]. The paramagnetic effect of enzyme-bound Mn(II) on the spin-lattice relaxation rates (T-1(1] of ligand nuclei (1H and 31P) in E.Mn(II).Glc and E.Mn(II).Glc6P complexes have been measured. The paramagnetic effect of Mn(II) on the proton relaxation rates of C1-H alpha, C1-H beta and C2-H beta of glucose in the E.Mn(II).Glc complex was measured at 270 MHz and 500 MHz. The temperature dependence of these rates was also studied in the range of 5-30 degrees C at 500 MHz. The ligand nuclear relaxation rates in E.Mn(II).Glc are field-dependent and the Arrhenius plot yields an activation energy (delta E) of 16.7-20.9 kJ/mol. Similar measurements have also been carried out on C1-H alpha, C1-H beta and C6-31P at 270 MHz (1H) and 202.5 MHz (31P) for the E.Mn(II).Glc6P complex. The temperature dependence of 31P relaxation rates in this complex was measured in the range 5-30 degrees C, which yielded delta E = 9.2 kJ/mol. The electron-nuclear dipolar correlation time (tau c), determined from the field-dependent measurements of proton relaxation rates in the E.Mn(II).Glc complex, is 0.22-1.27 ns. The distances determined between Mn(II) and C1-H of glucose and glucose 6-phosphate are approximately 1.1 nm and approximately 0.8 nm, respectively. These data, considered together with our recent results [Mehta, A., Jarori, G. K. & Kenkare, U. W. (1988) J. Biol. Chem. 263, 15492-15498], suggest that glucose and glucose 6-phosphate may bind to very nearly the same region of the enzyme. The structure of the binary Glc6P.Mn(II) complex has also been determined. The phosphoryl group of the sugar phosphate forms a first co-ordination complex with the cation. However, on the enzyme, the phosphoryl group is located at a distance of approximately 0.5-0.6 nm from the cation.  相似文献   

20.
Growth and contents of sodium (Na), potassium (K), calcium (Ca), magnesium (Mg), chloride (Cl), phosphorus (P) and sulphur (S) in shoot and root tissues of Carthamus tinctorius plants were measured at combinations of four nutrient solution osmotic potentials (s=0, -0.3, -0.6 and -0.9 MPa) induced by NaCl and CaCl treatments, three constant temperatures (T) ranging from 15 to 35°C and four abscisic acid (ABA) concentrations (0,10,50 and 100 mg L–1). Unstressed and stressed plants grown in optimal temperature conditions (25°C) maintained higher growth rates (dry mass production) than plants grown under low and high temperatures (15 and 35°C respectively). Shoot and root growth (dry mass production) were largely inhibited by salinity but the magnitude of growth inhibition was temperature dependent. Safflower plants respond to salinity stress by increases in Ca, Cl and to a lesser extent Na in their shoots and roots and by a decrease in the ratio of fresh to dry weight. The ratio of K/Na was decreased progressively on salinization. With stressed plants, ABA application reduced the toxicity of salt treatment, improved K uptake under salinity, effectively increased K/Na ratio and helped the plants to avoid Na toxicity and sometimes enhanced growth. The effect of ABA on the growth was more pronounced at optimum temperature (25°C). The association between the internal mineral element concentrations was largely affected by ABA application and temperature change but a wide fluctuation in response was noticed. The effects of single factors (s, T and ABA) on the growth and mineral contents were statistically significant. Also, bifactorial (s× T, s × ABA and T × ABA) and three factorial (s × T × ABA) interactions significantly affected the parameters. Further statistical treatment of the data (coefficient of determination 2) led to four important findings: (1) Salinity (s) was dominant in affecting Ca and Cl contents in both shoot and root as well as root Na content. (2) Temperature (T) had a dominant effect on growth, shoot K, Mg, P, S and root P, and S contents (3) The share of s × T × ABA interaction was dominant for root Na and Mg contents. (4) The single factors and their interactions had a dual role in their subsidiary effects.Abbreviations ABA abscisic acid - s osmotic potential - 2 coefficient of determination - F.wt fresh weight - d.m. dry matter - T temperature - MPa mega pascal - SAR sodium adsorption ratio - P phosphorus - S sulphur  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