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.
The nitric oxide synthase inhibitor N(G)-nitro-L-arginine methyl ester (l-NAME) increased vascular resistance (VR) 10% above baseline of 3.08+/-0.08 (n=11) mmHg/mL/min at 10 mg/kg and 20% above 3.05+/-0.08 (n=9) at 50 mg/kg in anesthetized toads (Bufo marinus). Blood pressure was unaffected by either dose of L-NAME. Blood flow decreased at the higher dose of L-NAME. L-arginine (300 mg/kg) reversed the effects of L-NAME on VR and blood flow in toads treated with 10 mg/kg but not with 50 mg/kg. Injection of 50 mg/kg L-NAME into empty-bladder toads produced a 10% decrease in water uptake, J(v), resulting in a J(v) of 1,267+/-11 cm(3)/cm(2)/s x 10(-7) (n=9) compared to 1,385+/-12 (n=8) for controls. Injection of 10 microg/kg angiotensin II (ANG II) increased J(v) 15% across the pelvic patch (J(v), cm(3)/cm(2)/s x 10(-7)), resulting in a J(v) of 1,723+/-12 cm(3)/cm(2)/s x 10(-7) (n=8) compared to 1,471+/-12 (n=8) for controls. It is hypothesized that during cutaneous drinking blood flow into the capillary bed of the pelvic patch is regulated by nitric oxide and ANG II.  相似文献   

2.
Dissociation of alpha beta and alpha gamma dimers of human hemoglobins (Hb) A and F into monomers was studied by alpha chain exchange (Shaeffer, J. R., McDonald, M. J., Turci, S. M., Dinda, D. M., and Bunn, H. F. (1984) J. Biol. Chem. 259, 14544-14547). Unlabeled carbonmonoxy-Hb A was incubated with trace amounts of preparatively purified, native, 3H-alpha subunits in 10 mM sodium phosphate, pH 7.0, at 25 degrees C. At appropriate times, free alpha monomers were separated from Hb A tetramers by anion exchange high performance liquid chromatography. Transfer of radioactivity from the alpha chain pool into Hb A was measured, yielding a first order dimer dissociation rate constant, k2 = (3.2 +/- 0.3) X 10(-3) h-1. The Arrhenius plot of k2 was linear between 7 and 37 degrees C, yielding an enthalpy of activation of 23 kcal/alpha beta dimer. As the chloride concentration was raised from 0 to 0.2 M, the dissociation rate increased 3-fold; with higher salt concentrations, however, the rate gradually returned to baseline. This rate was not altered by raising the pH from 6.5 to 7.2, but as pH was further raised to 8.4, kappa 2 increased about 3-fold. Hb F, which has an increased stability at alkaline pH, dissociated into alpha and gamma monomers 3 times more slowly than Hb A. Moreover, the dimer-monomer dissociation of Hb F was characterized by a significantly reduced pH dependence. These results demonstrate that both alpha beta and alpha gamma dimers of Hb A and Hb F dissociate reversibly into monomers under physiologic conditions. The differential pH dependence for dimer dissociation between Hb A and Hb F suggests that specific amino acid replacement at the alpha 1 gamma 1 interface confers increased resistance to alkaline denaturation.  相似文献   

3.
The reaction between carboxyhemoglobin and reduced microperoxidase (MP): Hb4(CO)4 + 4MP=Hb4 + 4MPCO, recently reported by us, has been further studied. By generating species Hb4(CO), Hb4(CO)2, and Hb(CO)3 in the stopped flow cuvette by the reaction of dithionite with the species of the general formula Hb4(O2)x(CO)y(x + y=4) in the presence of microperoxidase it has been possible to determine the stepwise CO dissociation rate constants l4, l3, l2, and l1. The overall CO dissociation rate constant l, which is the same in this system as l4, is not affected by 2,3-diphosphoglyceric acid. The activation energy of the reaction is 21,400 cal in 15-25 degrees range. The ratio deltal/deltapH is approximately 3 in 6.5 to 7.5 pH range. The kinetic data indicate that, compared to HbO2, the contribution to the cooperativity of the dissociation rate constants of carboxyhemoglobin is greatly reduced. The ligand-dependent differences in the reactions of Hb with CO, O2, and NO suggest that in the combination reactions the ligand plays an active role in the rate-limiting step.  相似文献   

4.
The kinetics of CO and NO recombination with the giant approximately 3600-kDa hexagonal bilayer hemoglobin of Lumbricus terrestris and its subunits, the approximately 200-kDa dodecamer of globin chains (3 x chains (I + II + III + IV] (see preceding paper (Vinogradov S.N., Sharma, P.K., Qabor, A.N., Wall, J.S., Westrick, J.A., Simmons, J.H., and Gill, S.J. (1991) J. Biol. Chem. 266, 13091-13096], the 50-kDa disulfide-bonded trimer (chains II-IV), the monomer (chain I), and the approximately 30-kDa linker (chains VA, VB, and VI), were measured following photolysis over time scales ranging from picosecond to millisecond. CO recombination at 436 nm subsequent to excitation (9 ns) at 532 nm showed three phases covering a 100-fold range for the Hb, dodecamer, trimer, and linker protein. The proportion of the fast phase was 0.1-0.2 for the trimer, dodecamer, and Hb. The relative rates and amplitudes of the phases were not affected by changes in CO concentration or excitation intensity. The monomer showed a single phase with a rate of 2 x 10(6) M-1 s-1. The second-order reaction with NO showed two rates. The faster rate was 90 x 10(6) M-1 s-1 and accounted for approximately 0.7 of the reaction for all species except the monomer, where it accounted for the full reaction. The slower rate was 15 x 10(6) M-1 s-1 for all species except the monomer.  相似文献   

5.
Laboratory bioassays were conducted to determine the relative susceptibility of eggs, 1st-, 3rd-, 5th-instar nymphs and adults of Rhodnius prolixus to one isolate of the entomopathogenic hyphomycete, Beauveria bassiana. Treatments consisted of directly spraying on insects of increasing doses of inoculum (3 x 10(2) to 3 x 10(5) conidia per cm2). Mortality due to all doses of conidia was very high in the five tested stages of the target insect. Experiments on eggs demonstrated that the fungal isolate was able to kill eggs before they hatched. Both time-mortality and dose-mortality responses showed that the susceptibility of R. prolixus varied according to its stage of development and increased with age. As a matter of fact, at the dose of 3 x 10(3) conidia per cm2, LD 50 varied between 11.2 days in 1st-instar nymphs and 6.4 days in both 5th-instar nymphs and adults. Comparison of LD50 permitted to estimate that 1st-instar nymphs were about 700-fold less susceptible than the two oldest stages.  相似文献   

6.
This study describes the kinetic behaviour and physicochemical aspects of an endogenous inhibitor of progesterone--receptor binding in trophoblast cytosol from day-12 embryos. The progesterone cytosol receptor was partially purified and isolated from the inhibitor as the 0--50%-satd. (NH4)2SO4 fraction. The inhibitory substance was shown to reside in the 50--70%-satd. (NH4)2SO4 fraction. Equilibration of the inhibitor preparation with the receptor fraction increased the Kapp.D of the ligand--receptor binding reaction in a concentration-dependent manner (26 +/- 3-fold increase in Kapp.D per mg of protein of the (NH4)2SO4 fraction, n = 16). However, the inhibitor did not alter the concentration of binding sites. Studies of other physicochemical aspects of the inhibitor showed it to be non-diffusible, excluded from Sephadex G-25, stable at 35 degrees C for 30 min, but irreversibly denatured at 70 degrees C for 30 min. The Stokes' radius was estimated by gel chromatography to be 2.8 +/- 0.11 nm (n = 5). Inhibitory activity was destroyed by HgCl2, suggesting that disulphide bridges play an essential role in the biological activity of this molecule. The inhibitor is a macromolecule which does not bind progesterone and differs from albumin. The kinetic mechanism by which the inhibitor enhanced Kapp.D was investigated by measuring association and dissociation rate constants and the energy of activation (Ea) for each reaction. The association rate (k+1) for progesterone and receptor was (1.3 +/- 0.2) x 10(4) M-1 . s-1 but declined to (0.4 +/- 0.1) x 10(4) M-1 . s-1 (n = 5) when exposed to the inhibitor (P less than 0.01). The dissociation rate (k-1) was (3.2 +/- 0.6) x 10(-5) s-1 for progesterone--receptor complex and was unchanged by the inhibitor. The Ea for the association of complex was 33.6 +/- 4.2 kJ/mol and was increased to 63.0 +/- 8.4 kJ/mol by the inhibitor (P less than 0.05). The Ea of dissociation was unaltered. Thus, an inhibitor is present in trophoblast cytosol which specifically enhances Kapp.D without altering availability of binding sites. The mode of action of inhibitor is to increase the energy of activation for association of complex without influencing the dissociation reaction.  相似文献   

7.
The metabolic response of mammary carcinoma in the C3H mouse to photodynamic therapy (PDT) was measured using in vivo 31P nuclear magnetic resonance (31P-NMR) spectroscopy and pH microelectrodes. Twenty-four hours after administration of Photofrin II (12.5 mg/kg), the tumor was subjected to photoactivation using an argon dye laser. Optical treatment doses were 200, 400, and 600 J/cm2 and corresponded to the following tumor control doses: TCD10/30, TCD50/30, and TCD90/30, respectively. In vivo 31P-NMR spectra and pH micro-electrode measurements were obtained prior to treatment and at 4, 24, 48, and 72 h and 1 week post-treatment. The data revealed a significant (P less than 0.0002) alkalosis as indicated by the pH measured by NMR compared to pH measured by microelectrodes at all treatment levels and time points. Spectral differences between treatment groups were apparent as early as 4 h after treatment. The ratio of beta-nucleoside triphosphate to inorganic phosphate at 4 h after treatment was significantly (P less than 0.01) smaller for 600 J/cm2 treatment than for 200 J/cm2 treatment. At curative (600 J/cm2) levels, from 48 h on, no phosphate resonances were detected in the spectra. The pH measured by NMR transiently decreased from pretreatment levels after 200 and 400 J/cm2 treatment (P less than 0.002, P less than 0.009, respectively), while no change in pH from pretreatment values was found after 600 J/cm2 treatment. The data suggest that the early metabolic response of mammary carcinoma to PDT, as indicated by 31P-NMR spectroscopy, is dose dependent, and may be a sensitive indicator of biological outcome to treatment.  相似文献   

8.
We reported increased water permeability and a low urea reflection coefficient in Xenopus oocytes expressing urea transporter UT-B (former name UT3), suggesting that water and urea share a common aqueous pathway (Yang, B., and Verkman, A. S. (1998) J. Biol. Chem. 273, 9369-9372). Although increased water permeability was confirmed in the Xenopus oocyte expression system, it has been argued (Sidoux-Walter, F., Lucien, N., Olives, B., Gobin, R., Rousselet, G., Kamsteeg, E. J., Ripoche, P., Deen, P. M., Cartron, J. P., and Bailly, P. (1999) J. Biol. Chem. 274, 30228-30235) that UT-B does not transport water when expressed at normal levels in mammalian cells such as erythrocytes. To quantify UT-B-mediated water transport, we generated double knockout mice lacking UT-B and the major erythrocyte water channel, aquaporin-1 (AQP1). The mice had reduced survival, retarded growth, and defective urinary concentrating ability. However, erythrocyte size and morphology were not affected. Stopped-flow light scattering measurements indicated erythrocyte osmotic water permeabilities (in cm/s x 0.01, 10 degrees C): 2.1 +/- 0.2 (wild-type mice), 2.1 +/- 0.05 (UT-B null), 0.19 +/- 0.02 (AQP1 null), and 0.045 +/- 0.009 (AQP1/UT-B null). The low water permeability found in AQP1/UT-B null erythrocytes was also seen after HgCl(2) treatment of UT-B null erythrocytes or phloretin treatment of AQP1 null erythrocytes. The apparent activation energy for UT-B-mediated water transport was low, <2 kcal/mol. Estimating 14,000 UT-B molecules per mouse erythrocyte, the UT-B-dependent P(f) of 0.15 x 10(-4) cm/s indicated a substantial single channel water permeability of UT-B of 7.5 x 10(-14) cm(3)/s, similar to that of AQP1. These results provide direct functional evidence for UT-B-facilitated water transport in erythrocytes and suggest that urea traverses an aqueous pore in the UT-B protein.  相似文献   

9.
Oxygen transport properties of blood in two different bovine breeds   总被引:1,自引:0,他引:1  
1. The whole oxygen dissociation curve of oxyhemoglobin has been determined in double-muscled cattle of the Belgian White Blue breed and in Friesian cattle of different body weight. 2. In calves, P50 values are low and DPG level is high (4-20 mumol/g Hb). 3. P50 values of 25 +/- 1.4 mm Hg (mean +/- SD) and a level of DPG less than 1.5 mumol/g Hb have been found in animals weighing more than 80 kg. 4. Effects of temperature and pH on the oxygen dissociation curve have been measured at all levels of saturation. The temperature coefficient (dlog P50/dT) and the Bohr effect expressed as dlog P50/dpH were 0.017 and -0.40, respectively. 5. Hematocrit, hemoglobin concentrations and oxygen capacity of hemoglobin have been measured. 6. No difference between both breeds has been observed. 7. These data can be used to correct measured values of oxygen tension for temperature and pH and to measure oxygen content of blood in cattle.  相似文献   

10.
The analytical method described in the preceding article was applied to spectrophotometric Ca2+-titrations of the metallochromic indicator arsenazo III (Ar). At various reactant concentrations it was determined that Ar forms 1:1,1:2 and 2 : 1 complexes with calcium. The equilibrium constants and extinction coefficients at 602 nm were determined. Corrected to zero ionic strength at 293 K and pH 7.0, the reactions Ca + Ar = CaAr, CaAr + Ar = CaAr2 and CaAr + Ca = Ca2Ar are associated with dissociation equilibrium constants k(11) = 1.6 x 10(-6)M, K12 = 3.2 x 10(-4)M and K21 = 5.8 x 10(-3)M. respectively. The extinction coefficient of unbound indicator is (602) = 9.6 (+/-0.3) x 10(3) cm(-1) M(-1). Arscnazo III complexes with monovalent ions like Na+ and K+ : at zero ionic strength, the dissociation constant of the Na+-Ar complex is about 0.1 M.  相似文献   

11.
血红蛋白携氧-释氧动力学研究   总被引:2,自引:0,他引:2  
Jiang C  Wang X  Gao W  Peng WY  Xie JX  Li YJ 《生理学报》2008,60(1):83-89
本文研究了鸡、家兔、鲤鱼、蟾蜍4种实验动物血红蛋白(hemoglobin,Hb)携氧-释氧动力学过程,初步建立Hb携氧-释氧动力学研究方法,并探讨Hb携氧-释氧动力学过程与动物生存环境之间的关系.结果显示:4种动物Hb携氧动力学曲线均呈"S"形曲线特征,与传统的Hb氧解离曲线(oxygen dissociation curve,ODC)相似;同时不同动物Hb携氧-释氧动力学曲线也有各自特点,如鸡Hb释氧时间长达(1 411±6)S;在Hb携氧.释氧曲线I阶段,鲤鱼上升斜率远大于家兔等.提示Hb携氧-释氧动力学曲线可反映不同动物Hb携氧效率的差异.与传统ODC参数P50相对应,由动力学曲线可得到Hb携氧动力学参数T50°T50是Hb达到50%氧饱和度所需时间,可直观反映Hb携氧效率的差异.4种实验动物Hb均有较稳定的T50,从大到小依次为:鸡、家兔、鲤鱼和蟾蜍.对Hb携氧动力学曲线与ODC综合分析,可得到Hb携氧效能参数E50,表示Hb达到50%氧饱和度所用时间与环境氧分压之间的关系,即E50(50% Sat,Xeo2,yr).E50有可能成为全面评价Hb携氧效能的综合指标.  相似文献   

12.
The diffusion coefficient (D) of 2,3-bisphosphoglycerate (DPG) was measured using pulsed-field gradient (PFG)-31P nuclear magnetic resonance spectroscopy in solutions containing 2.7-5.0 mM hemoglobin (Hb) and a range of DPG concentrations. The dependence of the measured values of D on the fraction of the total DPG in the sample that is bound to Hb enabled the estimation of the dissociation constants (Kd) of complexes of DPG with carbonmonoxygenated, oxygenated, and deoxygenated Hb; the values of Kd (mM), measured at 25 degrees C, pH 6.9 and in 100 mM bis Tris/50 mM KCl, were 1.98 +/- 0.26, 1.8 +/- 0.5 and 0.39 +/- 0.26, respectively. In intact erythrocytes the apparent diffusion coefficient, Dapp, of DPG was larger in oxygenated and carbonmonoxygenated cells (6.17 +/- 0.20 x 10(-11) m2s-1) than in deoxygenated cells (4.10 +/- 0.23 x 10(-11) m2s-1). Changes in intracellular DPG concentration (5-55 mM) in erythrocytes, brought about by incubation in a medium containing inosine and pyruvate, did not result in significant changes in the value of Dapp; this result supports the hypothesis that DPG binds to other sites in the erythrocyte. Monte Carlo simulations of diffusion in biconcave discs were used to test the adequacy of the values of Kd estimated in solution to describe the binding of DPG to Hb in oxygenated and deoxygenated erythrocytes. The results of the simulations implied that the value of Kd estimated for deoxygenated Hb-DPG was greater than expected from the experiments involving intact erythrocytes. This difference is surmised to be at least partly due to the difficulty of measuring D at low-ligand concentrations. Notwithstanding this shortcoming, the PFG method appears to be suitable for probing interactions between macromolecules and ligands when the Kd is in the millimolar range. It is one of the few techniques available in which these interactions can be studied in intact cells. In addition, the Monte Carlo simulations of the diffusion experiments highlighted important differences between theory and experiment relating to the nature of molecular motion inside the cells.  相似文献   

13.
Shan D  Han E  Xue H  Cosnier S 《Biomacromolecules》2007,8(10):3041-3046
A highly stable biological film was formed on the functional glassy carbon electrode (GCE) via step-by-step self-assembly of chitosan (CHT), laponite, and hemoglobin (Hb). Cyclic voltammetry (CV) of the Hb/laponite/CHT/GCE showed a pair of stable and quasi-reversible peaks for the Hb-Fe(III)/Fe(II) redox couple at about -0.035 V versus a saturated calomel electrode in pH 6.0 phosphate buffer at a scan rate of 0.1 V s(-1). The electrochemical reaction of Hb entrapped on the laponite/CHT self-assembled film exhibited a surface-controlled electrode process. The formal potential of the Hb-heme-Fe(III)/Fe(II) couple varied linearly with the increase of pH over the range of 3.0-8.0 with a slope of -63 mV pH(-1), which implied that an electron transfer was accompanied by single-proton transfer in the electrochemical reaction. The position of the Soret absorption band of this self-assembled Hb/laponite/CHT film suggested that the entrapped Hb kept its secondary structure similar to its native state. The self-assembled film showed excellent long-term stability, the CV peak potentials kept in the same positions, and the cathodic peak currents retained 90% of their values after 60 days. The film was used as a biological catalyst to catalyze the reduction of hydrogen peroxide. The electrocatalytic response showed a linear dependence on the H2O2 concentration ranging widely from 6.2 x 10(-6) to 2.55 x 10(-3) M with a detection limit of 6.2 x 10(-6) M at 3 sigma.  相似文献   

14.
The reaction between antiplasmin (A) and a low-molecular-weight form of plasmin (P) proceeds in at least two steps: a fast reversible second-order reaction followed by a slower irreversible first-order transition, and may be represented by: P +A k1 in equilibrium k-1 PA k2 leads to PA'. The low-Mr plasmin, which is obtained by limited elastase digestion, is composed of an intact B chain and a small A chain lacking the lysine-binding sites. The k1 of the reaction is (6.5 +/- 0.5) x 10(5) M-1 s-1 which is 30--60 times smaller than that for normal plasmin and antiplasmin. The dissociation constant of the first step is 1.9 x 10(-9) M which is 10 times higher than for normal plasmin and antiplasmin. The rate constant of the second step is (4.2 +/- 0.2) x 10(-3) s-1 for both normal and low-Mr plasmin. Low Mr plasmin which has substrate bound to its active site does not react or reacts only very slowly with antiplasmin. The reaction rate, however, is only slightly influenced by 6-aminohexanoic acid in concentrations up to 1 mM which decrease the reaction rate of normal plasmin approximately 50-fold. The findings further indicate that the lysine-binding site(s) of plasmin are of great importance for the rate of its reaction with antiplasmin.  相似文献   

15.
Three hemoglobins have been isolated from the symbiont-harboring gill of the bivalve mollusc Lucina pectinata. Oxyhemoglobin I (Hb I), which may be called sulfide-reactive hemoglobin, reacts with hydrogen sulfide to form ferric hemoglobin sulfide in a reaction that may proceed by nucleophilic displacement of bound superoxide anion by hydrosulfide anion. Hemoglobins II and II, called oxygen-reactive hemoglobins, remain oxygenated in the presence of hydrogen sulfide. Hemoglobin I is monomeric; Hb II and Hb III self-associate in a concentration-dependent manner and form a tetramer when mixed. Oxygen binding is not cooperative. Oxygen affinities are all nearly the same, P50 = 0.1 to 0.2 Torr, and are independent of pH. Combination of Hb I with oxygen is fast; k'on = (estimated) 100-200 x 10(6) M-1 s-1. Combination of Hb II and Hb III with oxygen is slow: k'on = 0.4 and 0.3 x 10(6) M-1 s-1, respectively. Dissociation of oxygen from Hb I is fast relative to myoglobin: koff = 61 s-1. Dissociation from Hb II and Hb III is slow: koff = 0.11 and 0.08 s-1, respectively. These large differences in rates of reaction together with differences in the reactions of carbon monoxide suggest differences in configuration of the distal heme pocket. The fast reactions of Hb I are comparable to those of hemoglobins that lack distal histidine residues. Slow dissociation of oxygen from Hb II and Hb III suggest that a distal residue may interact strongly with the bound ligand. We infer that Hb I may facilitate delivery of hydrogen sulfide to the chemoautotrophic bacterial symbiont and Hb II and Hb III may facilitate delivery of oxygen. The midpoint oxidation-reduction potential of the ferrous/ferric couple of Hb I, 103 +/- 8 mV, was independent of pH. Potentials of Hb II and Hb III were pH-dependent. At neutral pH all three hemoglobins have similar midpoint potentials. The rate constant for combination of ferric Hb I with hydrogen sulfide increases 3000-fold from pH 10.5 to 5.5, with apparent pK 7.0, suggesting that undissociated hydrogen sulfide is the attacking ligand. At the acid limit combination of ferric Hb I with hydrogen sulfide, k'on = 2.3 x 10(5) M-1 s-1, is 40-fold faster than combination with ferric Hb II or myoglobin.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

16.
A new Monte Carlo (MC) algorithm is proposed for simulating inextensible circular chains with finite twisting and bending rigidity. This new algorithm samples the relevant Riemann volume elements in a uniform manner, when the constraining potential vanishes. Simulations are performed for filaments comprising 170 subunits, each containing approximately 28 bp, which corresponds to a DNA length of 4770 bp. The bending rigidity is chosen to yield a persistence length, P = 500 A, and the intersubunit potential is taken to be a hard-cylinder potential with diameter d = 50 A. This value of d yields the same second virial coefficient as the electrostatic potential obtained by numerical solution of the Poisson-Boltzmann equation for 150 mM salt. Simulations are performed for unknotted circles and also for trefoil knotted circles using two different values of the torsional rigidity, C = (2.0 and 3.0) x 10(-19) dyne cm2. In the case of unknotted circles, the simulated supercoiling free energy varies practically quadratically with linking difference delta l. The simulated twist energy parameter ET, its slope dET/dT, and the mean reduced writhe <w>/delta l for C = 3 x 10(-19) dyne cm2 all agree well with recent simulations for unknotted circles using the polygon-folding algorithm with identical P, d, and C. The simulated ET vs. delta l data for C = 2.0 x 10(-19) dyne cm2 agree rather well with recent experimental data for p30 delta DNA (4752 bp), for which the torsional rigidity, C = 2.07 x 10(-19) dyne cm2, was independently measured. The experimental data for p30 delta are enormously more likely to have arisen from C = 2.0 x 10(-19) than from C = 3.0 x 10(-19) dyne cm2. Serious problems with the reported experimental assessments of ET for pBR322 and their comparison with simulated data are noted. In the case of a trefoil knotted DNA, the simulated value, (ET)tre, exceeds that of the unknotted DNA, (ET)unk, by approximately equal to 1.40-fold at magnitude of delta l = 1.0, but declines to a plateau about 1.09-fold larger than (ET)unk when magnitude of delta l > or = 15. Although the predicted ratio, (ET)tre/(ET)unk approximately equal to 1.40, agrees fairly well with recent experimental measurements on a 5600-bp DNA, the individual measured ET values, like some of those reported for pBR322, are so large that they cannot be simulated using P = 500 A, d = 50 A, and any previous experimental estimate of C.  相似文献   

17.
A double mixing stopped-flow method is described for studying the dimer-tetramer equilibria of oxyhemoglobins and the kinetics of association of unliganded dimers. The three hemoglobins studied were: Hb Carp, Hb A, and Hb Rothschild (Trp beta 37 (C3)----Arg). The new method reproduces the data obtained for oxyHb A by other established methods. In agreement with previous studies, the new method indicates little, if any, dissociation of oxyHb carp into dimers even in 2 M urea solutions (0.1 M Bis-Tris pH 7.0). OxyHb Rothschild, on the other hand, is extensively dissociated into dimers (K(Hb4L4 in equilibrium with 2Hb2) = 37.3 x 10(-6) M) and the rate constant for the association of deoxy dimers of Hb Rothschild is about one-tenth of the value for Hb A indicating that the deoxy tetramer of Hb Rothschild is at least 10 times more dissociated into dimers than deoxyHb A.  相似文献   

18.
To define aspects of lipid composition and bilayer asymmetry critical to barrier function, we examined the permeabilities of liposomes that model individual leaflets of the apical membrane of a barrier epithelium, Madin-Darby canine kidney type 1 cells. Using published lipid compositions we prepared exofacial liposomes containing phosphatidylcholine, sphingomyelin, glycosphingolipids, and cholesterol; and cytoplasmic liposomes containing phosphatidylethanolamine, phosphatidylserine, and cholesterol. The osmotic permeability of cytoplasmic liposomes to water (P(f)), solutes, and NH(3) was 18-90-fold higher than for the exofacial liposomes (P(f(ex)) = 2.4 +/- 0.4 x 10(-4) cm/s, P(f(cy)) = 4.4 +/- 0.3 x 10(-3) cm/s; P(glycerol(ex)) = 2.5 +/- 0.3 x 10(-8) cm/s, P(glycerol(cy)) = 2.2 +/- 0.02 x 10(-6) cm/s; P(NH3(ex)) = 0. 13 +/- 0.4 x 10(-4) cm/s, P(NH3(cy)) = 7.9 +/- 1.0 x 10(-3) cm/s). By contrast, the apparent proton permeability of exofacial liposomes was 4-fold higher than cytoplasmic liposomes (P(H+(ex)) = 1.1 +/- 0. 1 x 10(-2) cm/s, P(H+(cy)) = 2.7 +/- 0.6 x 10(-3) cm/s). By adding single leaflet permeabilities, we calculated a theoretical P(f) for a Madin-Darby canine kidney apical membrane of 4.6 x 10(-4) cm/s, which compares favorably with experimentally determined values. In exofacial liposomes lacking glycosphingolipids or sphingomyelin, permeabilities were 2-7-fold higher, indicating that both species play a role in barrier function. Removal of cholesterol resulted in 40-280-fold increases in permeability. We conclude: 1) that we have reconstituted the biophysical properties of a barrier membrane, 2) that the barrier resides in the exofacial leaflet, 3) that both sphingomyelin and glycosphingolipids play a role in reducing membrane permeability but that there is an absolute requirement for cholesterol to mediate this effect, 4) that these results further validate the hypothesis that each leaflet offers an independent resistance to permeation, and 5) that proton permeation was enhanced by sphingolipid/cholesterol interactions.  相似文献   

19.
Phytophthora ramorum, recently found in the US, is causing concern for hardwood forests and the nursery industry. In an effort to identify some of the environmental limitations to growth and sporulation we undertook a laboratory study of four US and three European (EU) isolates. On V8 media, isolates grew when incubated at 2-28 C and produced chlamydospores at 8-28 C. Sporangia were produced at all temperatures tested: 10-30 C for US isolates and 6-26 C for EU isolates. Optimal temperatures were 16-26 C for growth, 14-26 C for chlamydospore production and 16-22 C for sporangia production. US isolates grew less and produced fewer spores when exposed to increasing doses of near-UV radiation (50-300 microW/cm(2)) and visible radiation (250-1500 microW/cm(2)). EU isolates were exposed to 300 microW/cm(2) near-UV only, which significantly reduced growth of one of three isolates and had no significant effect on spore production. In our studies P. ramorum tolerated a broad range of temperature and light conditions, which suggests that it is capable of establishment in a wide geographic area.  相似文献   

20.
The structure and motion of phospholipids in human plasma lipoproteins have been studied by using 31P NMR. Lateral diffusion coefficients, DT, obtained from the viscosity dependence of the 31P NMR line widths, were obtained for very low density lipoprotein (VLDL), low-density lipoprotein (LDL), high-density lipoproteins (HDL2, HDL3), and egg PC/TO microemulsions at 25 degrees C, for VLDL at 40 degrees C, and for LDL at 45 degrees C. At 25 degrees C, the rate of lateral diffusion in LDL (DT = 1.4 x 10(-9) cm2/s) is an order of magnitude slower than in the HDLs (DT = 2 x 10(-8) cm2/s). At 45 degrees C, DT for LDL increases to 1.1 x 10(-8) cm2/s. In contrast, DT for VLDL increases only slightly going from 25 to 40 degrees C. The large increase in diffusion rate observed in LDL occurs over the same temperature range as the smectic to disordered phase transition of the core cholesteryl esters, and provides evidence for direct interactions between the monolayer and core. In order to prove the orientation and/or order of the phospholipid head-group, estimates of the residual chemical shift anistropy, delta sigma, have been obtained for all the lipoproteins and the microemulsions from the viscosity and field dependence of the 31P NMR line widths. For VLDL and LDL, the anisotropy is 47-50 ppm at 25 degrees C, in agreement with data from phospholipid bilayers. For the HDLs, however, significantly larger values of 69-75 ppm (HDL2) and greater than 120 ppm (HDL3) were obtained.(ABSTRACT TRUNCATED AT 250 WORDS)  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