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 46 毫秒
1.
The half-lives of elimination (T12) of 131I-RGG from the body of normal A or Balb/c animals was much longer than the T12 of SJL mice. At all ages, the T12 of normal hybrids (A × SJL, SJL × A, Balb/c × SJL) was similar to or longer than that of the A or Balb/c parents. Thus, in terms of the T12 of normal animals, the SJL responsiveness to 131I-RGG appeared to be a recessive trait. Tolerance could be induced in newborn animals and, in terms of T12, the degree of unresponsiveness at the age of 6 weeks, was the same in A, Balb/c, A × SJL, and Balb/c × SJL animals but was much shorter in SJL mice. Thus, in neonatally induced tolerance, the duration of tolerance was recessive for the SJL type. The average Tbuilt12 after tolerance induction in 3-week-old hybrids (A × SJL, SJL × A, Balb/c × SJL) was similar to that of the A or Balb/c parent, but by the 8th and 12th week it approached the average T12 of the SJL parent. Comparing 8-week-old hybrids, the average T12 was longest in A × SJL hybrids and identical in SJL × A and Balb/c × SJL mice. An examination of T12 distribution in various 8- and 12-week-old crosses and backcrosses revealed a fairly large proportion of individuals with a T12 which was intermediate between SJL and the other parent. There was a tendency for this number to decrease in 12 weeks as compared to 8-week-old mice. In 8-week-old mice, the number of animals with intermediate Tbuilt12 was smallest when SJL was the maternal animal [(SJL × A); SJL × (A × SJL); SJL × (SJL × A)]. There was no link between T12 of tolerant animals and either the immunoglobulin allotype (MuAl/MuA2) or the C5 eniotype (MuB1 positive/MuB1 negative).  相似文献   

2.
Mice, rendered tolerant to rabbit gamma globulin (RGG), were immunized with RGG or with dinitrophenylated RGG (DNP40-RGG), incorporated in adjuvant. The resulting response was evaluated in terms of the half-life of trace labeled RGG (131I-RGG). An antibody response against the tolerance inducing macromolecule could be elicited with DNP40-RGG, but not with RGG. Reconstitution experiments revealed that thymus derived (T) cells from tolerant donors could cooperate with bone marrow cells from normal donors in the response elicited by DNP40-RGG, but could not effectively cooperate with bone marrow derived (B) cells from tolerant donors. Such B cells could cooperate with T cells from normal donors. The relative difference between native and chemically modified proteins played an important role in this tolerance circumvention, since analogous experiments with human instead of rabbit gamma globulin did not result in an effective response to determinants of the tolerance-inducing proteins. It was suggested that the number of effectively immunogenic determinants on DNP40-RGG was low in B and in T cells of animals tolerant to RGG and that the probability of effective cooperation was consequently extremely low. If one of the two cell types came from a normal animal and thus could respond to a large number of determinants, the probability of effective cooperation increased so as to reveal the responsiveness of the “tolerant” cell population. There was no indication that the responsiveness of the tolerant T cell population was directed against tolerance-inducing determinants.  相似文献   

3.
4.
The longitudinal relaxation rate (1T1p) of water protons was studied in solutions of Mn(II)-concanavalin A at a number of frequencies. These relaxation rates were lowered in the presence of a variety of saccharides which have affinities for concanavalin A which range over two orders of magnitude. A good correlation was found in which saccharides which bind tightly have the greatest effect and saccharides which bind weakly or not at all have little effect on the 1T1p values. The temperature dependence of the proton relaxation rates showed that the lowering of these rates in the presence of saccharides was most likely due to a change in the exchange rate of solvent interacting with protein-bound Mn(II), 1Tm.An analysis of the temperature and frequency dependence of the 1T1p and 1T2p (transverse) solvent proton relaxation rates resulted in evaluation of a number of parameters for solvent water molecules interacting in the first coordination sphere of Mn(II) bound to concanavalin A. The ratio of the number of water molecules (q) to the Mn(II)-proton distance (r) obtained from a computer fit of the data over a limited temperature range is in accord with the findings of Koenig et al. ((1973) Proc. Nat. Acad. Sci.70, 475) and Meirovitch and Kalb ((1973) Biochim. Biophys. Acta303, 258). However, our studies of 1T1p and 1T2p of water over a more extensive temperature range are best fit with the following conclusions: at low temperatures (<20 °C), the data are consistent with an outer-sphere relaxation process. At higher temperatures (> 30 °C), the water molecule in the inner coordination sphere of the bound Mn(II) begins exchanging more rapidly and contributes to the relaxation processes (1T1p and 1T2p). The relaxation time of protons in the inner coordination shell, T1M, contributes over the entire temperature range and produces a frequency dependence in the relaxivity data from 6 to 100 MHz since the contributions to the correlation times are in the range 10?9-10?8 sec.  相似文献   

5.
The hydration properties of Escherichia coli lipids (phosphatidylglycerol, phosphatidylethanolamine) and synthetic 1,2-dioleoyl-sn-glycero-3-phosphocholine in H2O/2H2O mixtures (9:1, v/v) were investigated with 2H-NMR. Comparison of the 2H2O spin lattice relaxation time (T1) as a function of the water content revealed a remarkable quantitative similarity of all three lipid-H2O systems. Two distinct hydration regions could be discerned in the T1 relaxation time profile. (1) A minimum of 11–16 water molecules was needed to form a primary hydration shell, characterized by an average relaxation time of T1 ≈ 90 ms. (2) Additional water was found to be in exchange with the primary hydration shell. The exchange process could be described in terms of a two-site exchange model, assuming rapid exchange between bulk water with T1 = 500 ms and hydration water with T1 = 80–120 ms. Analysis of the linewidth and the residual quadrupole splitting (at low water content) confirmed the size of the primary hydration layer. However, each lipid-water system exhibited a somewhat different linewidth behavior, and a detailed molecular interpretation appeared to be preposterous.  相似文献   

6.
7.
The spin labels, 5-nitroxide stearic acid and 16-nitroxide stearic acid were incorporated into whole sciatic nerves dissected from normal, quaking, jimpy and trembler mice. With 5-nitroxide stearic acid, we have studied the thermal variation of the maximal apparent coupling constant (T6) between 0°C and 50°C. Within this range of temperatures, we obtained identical values of 2 T6 for nerves from normal and jimpy mice, whereas 2 T6 was smaller for nerves from quaking and trembler mice. With 16-nitroxide stearic acid, composite spectra were recorded, particularly in the high-field range. A line characteristic of myelin was clearly observed in the spectra of nerves from normal and jimpy mice; its intensity was somewhat less in nerves from quaking mice and much less in spectra from trembler mice. A shoulder in the principal highfield line of the spectrum is modified only with nerves from jimpy mice.The results agree well with those obtained by electron microscopy, which reveal normal myelination in nerves from jimpy mice, a slight modification of the myelin from those of quaking mice and a practically complete demyelination in peripheral nerves from trembler mice. However, the structure of the nerves of jimpy mice also seems to be modified at an, as yet, undetermined level.  相似文献   

8.
The effect of cold exposure caused by shearing on serum thyroid hormone (TH) concentrations in sheep kept at an ambient temperature of 8.5°C was studied. While the deep body temperature fell to the lowest level 4 h after shearing the concentration of triiodothyronine (T3) increased to a peak value at that time. Thyroxine (T4) and metabolically inactive reverse triiodothyronine (rT3) levels reached their peak value after 24 h. The T3T4 ratio reached a maximum at about 4 h and rT3T4 and rT3T3 ratios rose to maximum values about 24 h after shearing. This sequence of events suggest a biphasic response to cold—an immediate secretion of TH from the thyroid gland, followed by adaptive alteration in T3 and rT3 generation in the extrathyroidal tissues.  相似文献   

9.
Respiration (O), ammonium (NH4), phosphate (PO4), total nitrogen (NT) and phosphorus (PT) excretions were measured on mixed zooplankton during 3-, 6-, 9-, 12-, 21-, and 24-h incubation periods at 20–23 C. The excretion rates of PO4, NT. and PT decrease during a 21-h period, while rates of respiration and excretion of NH{IN4} are constant. The percentage of inorganic nitrogen excreted increases regularly from 3 h (30–40% of total nitrogen) to 21 h (70–80%) and it could be either due to a bacterial activity which was measured or to a decrease with time of organic nitrogen excreted because of starvation. ONT, OPO4, OPT, and NH4PO4 ratios increase during the first 9 h of incubation; the percentage of inorganic phosphorus excreted is higher at the very beginning and then remains constant from 6 to 24 h. ONH4 and NTPT ratios are constant during a 24-h term, which makes them useful metabolic indexes.  相似文献   

10.
Cytochrome b5 was extracted and purified from beef liver by a detergent method (cytochrome d-b5). The hydrophilic moiety which carries the heme group (cytochrome t-b5) was prepared by trypsin action upon pure cytochrome d-b5.Single-shelled lecithin liposomes form complexes with cytochromes d-b5 up to a molar ratio of one protein for 35 phospholipids. The lipid-protein complexes were isolated by gel filtration on Sepharose 4B. They are hollow vesicles in which [3H]-glucose can be trapped. Their diameter is greater than that of the initial liposomes.Cytochrome t-b5 does not interact with the vesicles. These results show that the hydrophobic tail is necessary for the binding and that the hydrophilic part of the protein is located on the outer face of the vesicles. This asymmetry is also proved by the action of reducing agents.Experiments with saturated phosphatidylcholines show that the protein interacts with the lipids both below the transition temperature TM. i.e. when the aliphatic chains are in a crystalline state, and above TM, when the alipathic chain are in a fluid state.1H NMR spectra show that even at the maximum cytochrome d-b5 concentration the presence of the proteins does not markedly change the dynamics to the phospholipid molecules. An asymmetric single-shelled vesicle structure is proposed for the complex.  相似文献   

11.
12.
In the redheaded bunting Emberiza bruniceps, thyroidectomy inhibited premigratory fattening and nocturnal restlessness—two characteristics of avian migration—observed in caged birds during the premigratory period (March/April). Thyroxine (T4) and triiodothyronine (T3) administration in thyroidectomized birds stimulated locomotor activity and restored the loss in body weight. Annual variations in circulating thyroid hormone concentrations revealed a significant rise in T3T4 ratio prior to spring migration in both years studied. This increase in circulating T3T4 ratio may be associated with the development of migratory disposition in this bird. There was no increase in circulating T3T4 ratio prior to autumnal migration, however, plasma T4 increased significantly. Different thyroidal mechanisms are most likely involved in spring and fall migratory periods. While T3 remained low throughout, apart from the characteristic spring rise, high T4 levels in E. bruniceps were associated with periods of reproduction and molting, the latter coinciding partly with autumnal migration.  相似文献   

13.
Binding of the chromogenic ligand p-nitrophenyl α-d-mannopyranoside to concanavalin A was studied in a stopped-flow spectrometer. Formation of the protein-ligand complex could be represented as a simple one-step process. No kinetic evidence could be obtained for a ligand-induced change in the conformation of concanavalin A, although the existence of such a conformational change was not excluded. The entire change in absorbance produced on ligand binding occurred in the monophasic process monitored in the stopped-flow spectrometer. The value of the apparent second-order rate constant (ka) for complex formation (ka = 54,000 s?1m? at 25 °C, pH 5.0, Γ/2 0.5) was independent of the protein concentration when the protein was in the range of 233–831 μm in combining sites and in excess of the ligand. The apparent first-order rate constant (k?a) for dissociation of the complex was obtained from the rate constant for the decomposition of the complex upon the addition of excess methyl α-d-mannopyranoside (k?a = 6.2 s?1 at 25 °C, pH 5.0, Γ/2 0.5). The ratio ka?a (0.9 × 104m?1) was in reasonable agreement with value of 1.1 ± 0.1 × 104m?1 determined for the equilibrium constant for complex formation by ultraviolet difference spectrometry. Plots of ln(kaT) and ln(kaT) vs 1T were linear (T is temperature) and were used to evaluate activation parameters. The enthalpies of activation for formation and dissociation of the complex are 9.5 ± 0.3 and 16.8 ± 0.2 kcal/mol, respectively. The unitary entropies of activation for formation and dissociation of the complex are 2.8 ± 1.1 and 1.3 ± 0.7 entropy units, respectively. These entropy changes are much less than those usually associated with substantial changes in the conformation of proteins.  相似文献   

14.
We determine the kinetic parameters V and KT of lactose transport in Escherichia coli cells as a function of the electrical potential difference (Δψ) at pH 7.3 and ΔpH = 0. We report that transport occurs simultaneously via two components: a component which exhibits a high KT (larger than 10 mM) and whose contribution is independent of Δψ, a component which exhibits a low KT independent of Δψ (0.5 mM) but whose V increases drastically with increasing Δψ. We associate these components of lactose transport with facilitated diffusion and active transport, respectively. We analyze the dependence upon Δψ of KT and V of the active transport component in terms of a mathematical kinetic model developed by Geck and Heinz (Geck, P. and Heinz, E. (1976) Biochim. Biophys. Acta 443, 49–63). We show that within the framework of this model, the analysis of our data indicates that active transport of lactose takes place with a H+/lactose stoichiometry greater than 1, and that the lac carrier in the absence of bound solutes (lactose and proton(s)) is electrically neutral. On the other hand, our data relative to facilitated diffusion tend to indicate that lactose transport via this mechanism is accompanied by a H+/lactose stoichiometry smaller than that of active transport. We discuss various implications which result from the existence of H+/lactose stoichiometry different for active transport and facilitated diffusion.  相似文献   

15.
A complete titration of phosphatidic acid bilayer membranes was possible for the first time by the introduction of a new anaologue, 1,2-dihexadecyl-sn-glycerol-3-phosphoric acid, which has the advantage of a high chemical stability at extreme pH values. The synthesis of this phosphatidic acid is described and the phase transition behaviour in aqueous dispersions is compared with that of three ester phosphatidic acids; 1,2-dimyristoyl-sn-glycerol-3-phosphoric acid, 1,3-dimyristoylglycerol-2-phosphoric acid and 1,2-dipalmitoyl-sn-glycerol-3-phosphoric acid.The phase transition temperatures (Tt) of aqueous phosphatidic acid dispersions at different degrees of dissociation were measured using fluorescence spectroscopy and 90° light scattering. The Tt values are comparable to the melting points of the solid phosphatidic acids in the fully protonated states, but large differences exist for the charged states.The Tt vs. pH diagrams of the four phosphatidic acids are quite similar and of a characteristic shape. Increasing ionisation results in a maximum value for the transition temperatures at pH 3.5 (pK1). The regions between the first and the second pK of the phosphatidic acids are characterised by only small variations in the transition temperatures (extended plateau) in spite of the large changes occurring in the surface charge of the membranes. The slope of the plateau is very shallow with increasing ionisation. A further decrease in the H+ concentration results in an abrupt change of the transition temperature. The slope of the Tt vs. pH diagram beyond pK2 becomes very steep. This is the  相似文献   

16.
Using 13C cross-polarization NMR techniques, we have found that the effect of protein on the dynamics of the hydrocarbon interior of a series of biological membranes is to depress the intensity of motion on the nanosecond timescale (i.e., T1 becomes longer) and to enhance the intensity of motion on the timescale of tens of microseconds (i.e., T1p becomes shorter).  相似文献   

17.
A method is described to measure the oxygen diffusion-concentration product, Do[O2], at any locus that can be probed or labeled using nitroxide radicals. The method is based on the dependence of the spin-lattice relaxation time T1 of the spin label on the bimolecular collision rate with oxygen. Strong Heisenberg exchange between spin label and oxygen contributes directly to T1 of the spin label, while dipolar interactions are negligible. Both time-domain and continuous wave saturation methods for studying T1 are considered. The method has been applied to phospholipid liposomes using fatty acid spin labels. A discontinuity in Do[O2] at the main phase transition was observed.  相似文献   

18.
Phosphate transporter of bovine heart mitochondria was purified by solubilization of submitochondrial particles with octylglucoside and fractionation of the extract with ammonium sulfate. After reconstitution into liposomes the purified protein catalyzed phosphate transport which was sensitive to mersalyl and other SH reagents. Transport measured either as PiOH or PiPi exchange was proportional to protein concentration and time. The PiOH but not the PiPi exchange was stimulated several fold by valinomycin plus nigericin in the presence of K+. The reconstituted system provides a suitable assay during purification of the mitochondrial phosphate transporter.  相似文献   

19.
Synthesis and phase transition characteristics of aqueous dispersions of the homologous (12 : 0, 14 : 0, 16 : 0) diphosphatidylglycerols (cardiolipins) and phosphatidyldiacylglycerols are reported. Electron microscopy of the negatively stained aqueous dispersions reveals a characteristic lamellar structure suggesting that these phospholipid molecules are organized as bilayers in the aqueous dispersions. The phase transition temperature (Tm) and the enthalpy of transition (ΔH) increase monotonically with chain length in the cardiolipin and phosphatidyldiacylglycerol series; Tm for phosphatidyldiacylglycerol is higher than that for cardiolipin of the same chain-length. The transition temperatures for the enantiomeric sn-3,3- and sn-1,1-phosphatidyldiacylglycerol and for the diastereomeric, meso-sn-1,3-phosphatidyldiacylglycerol are approximately the same. The molar enthalpy for the transition of cardiolipin-NH4+ bilayers is approximately twice the value for the phosphatidylcholines of the same chain length, i.e., the molar enthalpy per acyl chain is approximately the same in the two systems. The transition temperatures for metal ion salts of C1 6-cardiolipin exhibit a biphasic dependence upon the unhydrated ionic radii, i.e. the highest Tm is observed for Ca2+- cardiolipin and decreases for the salts of ions with smaller and larger ionic radii than that of Ca2+. The lowest Tm is observed for Rb+-cardiolipin. Monovalent metal salts of cardiolipin exhibit two phase transitions. This effect may result from different conformational packing of the four acyl chains due to differences in metal-phosphate binding.  相似文献   

20.
Perturbations induced by melittin on the thermotropism of dimyristoyl-, dipalmitoyl-, distearoylphosphatidylcholine and natural sphingomyelin are investigated and rationalized from data obtained by fluorescence polarization, differential scanning calorimetry and Raman spectroscopy. Depending on the technique and / or experimental conditions used, the observed effects differ at the same lipid to protein molar ratio, due to partial binding of melittin. The binding is more efficient for tetrameric than for monomeric melittin, but in both cases its affinity is weaker for phosphatidylcholine dispersions in the gel phase than for sonicated vesicles. For temperatures T ? Tm efficient binding occurs whatever the initial state of the lipids is. One can summarize the effects induced by melittin on the transition temperature as follows: (i) No upward shift is observed on synthetic phosphatidylcholines when lipid degradation is avoided. This is achieved by using highly purified melittin, phospholipase inhibitors, and / or non-hydrolysable lipids. (ii) Melittin monomer does not change Tm. (iii) When melittin tetramer is stabilized, it decreases Tm by 10–15 deg. C. The transition broadens, and is finally abolished for Ri ? 2. Very similar results are found for natural sphingomyelin. Fluorescence polarization indicates similar changes in order and dynamics of the acyl chains for all lipid studied. For T ? Tm, fluorescence and Raman show that melittin decreases the amount of CH2 groups in ‘trans’ conformation and the intermolecular order of the chains. According to fluorescence data, there is an increase of the rigid-body orientational order at T ? Tm, while from Raman the positional intermolecular order decreases without significant change in the CH2 groups ‘trans’/‘gauche’ ratio.  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