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.
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   

2.
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   

3.
Binding of the structural protein soc to the head shell of bacteriophage T4   总被引:5,自引:0,他引:5  
Qβ plus strands with a 70 S ribosome bound to the coat cistron initiation site were used as template for Qβ replicase. Minus strand synthesis proceeded until the replicase reached the ribosome. The ribosome was removed and elongation was continued in a substrate-controlled, stepwise fashion. The nucleotide analog N4-hydroxyCMP was introduced into the positions complementary to the third and fourth nucleotides of the coat cistron. The minus strands were elongated to completion, purified and used as template for Qβ replicase. The final plus strand preparation consisted of four species, with the sequences -A-U-G-G- (wild type), -A-U-A-G- (mutant C3), -A-U-G-A- (mutant C4) and -A-U-A-A- (mutant C3C4) at the coat initiation site. The ribosome binding capacity of the mutant RNAs relative to wild type was <0.1 (C3), 3.2 (C4) and 0.3 (C3C4). The finding that mutant C3 no longer formed an initiation complex suggests that the interaction of the ribosome binding site with fMet-tRNA plays an essential role in the formation of the 70 S initiation complex. The fact that mutant C4 RNA bound more efficiently than wild type, and that mutant C3C4 RNA showed substantial ribosome binding capacity whereas the single mutant C3 did not, can be explained by assuming that an A residue following the A-U-G triplet interacts with a complementary U residue in the anticodon loop sequence. In the case of C3C4 this additional base-pair may offset the reduced codon-anticodon interaction resulting from the modification of the A-U-G codon.  相似文献   

4.
5.
Regulation of 25-hydroxyvitamin D-3 24-hydroxylase by 1,25-dihydroxyvitamin D-3 and synthetic human parathyroid hormone fragment 1–34 (PTH1–34) was investigated using a cloned monkey kidney cell line, JTC-12. Treatment of the cells with 1,25-dihydroxyvitamin D-3 markedly enhanced the conversion of [3H]-25-hydroxyvitamin D-3 into a more polar metabolite. The metabolite was identified as 24,25-dihydroxyvitamin D-3 by normal phase and reverse phase high-performance liquid chromatography and periodate oxidation. The 24-hydroxylae activity appeared to follow Michaelis-Menten kintics, and 1,25-dihydroxyvitamin D-3 treatment increased the Vmax of 24-hydroxylase from 33 to 95 pmol/h per 106 cells without affecting the apparent Km value of the enzyme (220 nM in control vs. 205 nM in 1,25-dihydroxyvitamin D-3 treated cells). The enzyme activity reached a maximum between 4 and 8 h of treatment with 1,25-dihydroxyvitamin D-3. The dose of 1,25-dihydroxyvitamin D-3 required to cause a half-maximal stimulation was about 3 · 10?10 M. The 1,25-dihydroxyvitamin D-3-induced increase in 24-hydroxylase was almost completely inhibited by the presence of 1 μM cycloheximide. Treatment of the cells with PTH1–34 caused a dose-dependent increase in cyclic AMP production. Half-maximal stimulation of cyclic AMP production was obtained at about 5 · 10?9 M PTH1–34. When 2.4 · 10?9 M PTH1–34 was added after 1,25-dihydroxyvitamin D-3 treatment, the 1,25-dihydroxyvitamin D-3-stimulated 24-hydroxylase was inhibited to 70.7 ± 2.9% of control. Higher concentrations of PTH1–34 caused less inhibition of the enzyme activity. When cyclic AMP was added instead of PTH1–34, the enzyme activity was also suppressed significantly. These results indicate that, in JTC-12 cells, 1,25-dihydroxyvitamin D-3 stimulates 24-hydroxylase in a dose- and time-dependent manner by increasing the Vmax of the enzyme through a mechanism dependent upon new protein synthesis, and suggest that PTH1–34 inhibits the 1,25-dihydroxyvitamin D-3-induced stimulation of 24-hydroxylase through its effect on cyclic AMP production.  相似文献   

6.
7.
8.
9.
α-Chymotrypsin, converted to the acetyl enzyme by the p-nitrophenyl esters of CH3COOH, CH2DCOOH, CHD2COOH, and CD3COOH, undergoes deacetylation at pH 7.6 (phosphate buffer) and 25°C with secondary isotope effects of k(CH3)k(CH2D) = 0.985 ± 0.006, k(CH3)k(CHD2) = 0.971 ± 0.010, and k(CH3)k(CD3) = 0.956 ± 0.008. These isotope effects obey the simple additivity rule (“Rule of the Geometric Mean”) to within 20 J/mol, corresponding to about 5–6% of the maximum isotope effect for carbonyl addition. Thus, to this level, the three hydrogenic sites of the acetyl group are not rendered distinct in their contributions to the overall isotope effect even in the chiral environment of the chymotrypsin active site.  相似文献   

10.
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   

11.
12.
The changes in polymer-solvent interactions that occur when native calf thymus DNA is dialyzed against Na2SO4 solutions of a given ionic strength and buffer concentration but of varying concentrations in methylmercuric hydroxide have been investigated with the help of solution density measurements at 25 °C and pH 6.8–7.0. From measurements executed under equilibrium dialysis conditions at the three salt levels 5 mm, 0.05 m, and 0.5 m Na2SO4 (m refers to molality) and in the presence of 5 mm cacodylic acid buffer, the density increments (???c2)μ0 for native calf thymus DNA were determined as a function of CH3HgOH concentration. (???c2)μ0 was found not to vary with organomercurial concentration, irrespective of the concentration of supporting electrolyte, until a certain CH3HgOH concentration level has been reached, viz., pM1 ? 3.5 (pM1 = ?log mCH3HgOH), beyond which (???c2)μ0 increases strongly with increasing concentration of CH3HgOH. As is shown by optical melting, (???c2)μ0 becomes a function of organomercurial concentration the moment DNA undergoes denaturation brought about by the complexing of CH3HgOH with the various N-binding sites of the base residues in the DNA double helix.Polymer-solvent interactions, expressed in terms of preferential water interactions (“net hydration”) and preferential salt interactions (“salt solvation”), were derived from the (???c2)μ0 data in combination with data obtained on the preferential interaction of CH3HgOH with denatured DNA and data on the partial specific volumes of all major solution components, gathered from density measurements on solutions with fixed concentrations of diffusible components. Evidence is presented which shows that denaturation in general decreases the net hydration while salt becomes preferentially associated with the polyelectrolyte. This process is further amplified by the interaction of CH3HgOH with denatured DNA: Methylmercurated DNA alters the redistribution of diffusible components at dialysis equilibrium to such an extent that in a formal sense large amounts of water are rejected from the immediate vicinity of the polymer. The molecular implications of these findings are explored. The results are further discussed in the light of previous findings where the methylmercury-induced denaturation of DNA had been studied with the help of buoyant density measurements in a Cs2SO4 density gradient and by velocity-sedimentation in a variety of sulfate media.  相似文献   

13.
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   

14.
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   

15.
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   

16.
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   

17.
Unidirectional fluxes of [14C]lactose by whole cells of Escherichia coli under highly energized and partially de-energized (in the presence of CN?) conditions are analyzed kinetically.When the cells are energized, the value for V influx is 0.45 ± 0.01 mM internal concentration increment/s and Kt is 0.26 ± 0.03 mM. At an external concentration of 0.61 mM the steady-state internal concentration is 0.25 M, reached after about 1h. The maximum steady-state concentration ratio is 2 · 103.The efflux process under these conditions is non-saturable, being linearly dependent upon internal concentration over the range 25–250 mM with a first-order rate constant of 8.8 ± 0.2 · 10?4 s?1.The transport in the presence of CN? is active, with a maximum concentration ratio (internal concentration/external concentration) of 104, and the uptake is mimicked by anoxia (< 70 ppm O2).The effects of CN? are to lower the V for influx and to change the efflux from a non-saturable to a saturable process with a value for Kt (60 mM) intermediate between that for energized efflux (> 250 mM) and influxe (0.3–0.6 mM), the latter value not changing appreciably. Partial de-energization thus affects both the influx and efflux processes.  相似文献   

18.
The hydrogen content of whole human breath (not concentrated or end-expiratory samples), useful as an index of carbohydrate malabsorption, can be determined accurately by gas chromatographic analysis using a thermal conductivity detector and, for the first time, and electronic peak integrator. Standard gas mixtures (H2Ar or preferably H2N2) give peak areas linearly related to hydrogen content. For repeatable results, it is necessary to pressurize samples in a 0.5-ml sample loop. Five minutes are required for the analysis of a sample with a printed report of the area of the H2 peak (arbitrary units) and concentration of H2 (parts per million). Standard samples can be stored in special impervious bags for up to 3 days without loss. Details of bag construction are given.  相似文献   

19.
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   

20.
Amiloride in nM to μM concentrations stimulates the short circuit current (Isc) of the toad urinary bladder by as much as 120% when applied in conjunction with apical Ca2+ and a divalent cation chelator. A significant decrease in transepithelial resistance (Rt) is observed simultaneously. This response is spontaneously reversible and its amplitude is dependent upon apical sodium concentrations. The stimulated Isc persisted when acetazolamide (1 mM) was introduced, HPO42? substituted for HCO3? or SO42? replaced Cl?. Consequently, the increase in Isc is not due to the change of Cl?, H+ or HCO3? flux. This behavior in a ‘tight’ epithelium may be related to the mechanism controlling apical sodium permeability.  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