首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Synthesis and phase transition chaaracteristics 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 (delta H) 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-NH+4 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 C16-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   

2.
The interaction of cardiolipin with Ca2+ was assessed by measuring the cardiolipin-mediated extraction of 45Ca2+ from an aqueous to an organic (methylene chloride) phase. Cardiolipin binds Ca2+ with high affinity [Kd(apparent)=0.70±0.17 μM (S.D.)]. Cation-cardiolipin interactions are selective. Interaction of cardiolipin with Ca2+ is insensitive to Na+, but is inhibited by divalent cations with Mn2+ >Zn2+ >Mg2+. In addition La3+ and Ruthenium red are particularly potent inhibitors of Ca2+ binding by cardiolipin. Cardiolipin-mediated extraction of Ca2+ into an aqueous phase is also inhibited by phosphatidylcholine. Inhibition of Ca2+-cardiolipin interaction by phosphatidylcholine (a phospholipid known to stabilize the bilayer conformation) may implicate inverted, non-bilayer lipid structures in the binding.  相似文献   

3.
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   

4.
Equilibrium and kinetics of thermal melting of yeast 5.8S ribosomal RNA in aqueous NaCl were investigated by differential thermal melting and temperature jump methods. Two peaks were observed in each of the melting curves at 1 mM-1 M Na+ and linearity between each melting temperature Tm and log[Na+] was found at [Na+> 10 mM. From the difference spectrum ratio, dA280dA260, the G-C content in the local structures was calculated to be 91 and 56%. The temperature jump to 70–85°C in aqueous 30 mM Na+ of the RNA solution induced first-order kinetics, from which the kinetically determined melting curve was calculated. The curve could be approximately described in a Gaussian form with a Tm which agrees well with the high Tm in the static melting curve at 30 mM Na+. The kinetic properties of the reaction indicated a double helix-coil transition. However, the temperature jump to 20–60°C did not induce monophasic kinetics. The kinetic amplitude of the slow component showed a Tm which corresponded to the low Tm in the static melting curve at 30 mM Na+. The slow relaxation had the characteristics of a double helix-to-coil transition. However, contributions from very fast processes including single strand unstacking, were most noticeable in the low temperature melting region of the static curve. The thermodynamic parameters of both transitions from double helix to coil were analysed in detail. Both activation energies for helix formation were negative, and the nucleation is thought to follow a process similar to that in oligonucleotides. Values of Tm and enthalpy change of both helix-coil transitions indicated the cloverleaf model as the most plausible one for some limited regions of yeast 5.8S RNA among the previously proposed models: burp gun, cloverleaf and Rubin's models.  相似文献   

5.
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   

6.
An aqueous dispersion of fully hydrated bovine sphingomyelin was studied using 14N-NMR spectroscopy. Spectra were obtained as a function of temperature over the range 15–80°C, in both the liquid crystal and gel phases. In the liquid crystal phase, powder pattern lineshapes were obtained, whose quadrupolar splitting slowly decreases with increasing temperature. The spectra are increasingly broadened as the temperature is lowered through the phase transition into the gel phase. The linewidths and the second moments of these spectra indicate that the onset of a broad phase transition occurs at approx. 35°C, in agreement with previous calorimetric and 31P-NMR measurements. There is no evidence from the lineshapes for an hexagonal phase in this system, and this conclusion is supported by X-ray diffraction measurements carried out on aqueous dispersions of sphingomyelin in both phases. Assuming that the static nitrogen quadrupole coupling constant is the same for both sphingomyelin and dipalmitoyl-l-α-phosphatidylcholine (DPPC), the decrease observed in the quadrupolar splitting of sphingomyelin compared to that of DPPC indicates that the orientational order of the choline headgroup in liquid crystalline sphingomyelin is not the same as that of its counterpart in DPPC. Preliminary relaxation time measurements of T1 and T2 are presented which suggest that there are also dynamic differences between sphingomyelin and DPPC in the choline headgroup.  相似文献   

7.
A tRNAPhe derivative carrying ethidium at position 37 in the anticodon loop has been used to study the effect of spermine on conformational transitions of the tRNA. As previously reported (Ehrenberg, M., Rigler, R. and Wintermeyer, W. (1979) Biochemistry 18, 4588–4599) in the tRNA derivative the ethidium is present in three states (T1–T3) characterized by different fluorescence decay rates. T-jump experiments show two transitions between the states, a fast one (relaxation time 10–100 ms) between T1 and T2, and a slow one (100–1000 ms) between T2 and T3. In the presence of spermine the fast transition shows a negative temperature coefficient indicating the existence of a preequilibrium with a negative reaction enthalpy. Spermine shifts the distribution of states towards T3, as does Mg2+, but the final ratio [T2][T1] obtained with spermine is higher than with Mg2+, which we tentatively interpret to mean that spermine stabilizes one particular conformation of the anticodon loop.  相似文献   

8.
The mode of interaction of aqueous dispersions of phospholipid vesicles is investigated. The vesicles (average diameter 950 Å) are prepared from total lipid extracts of Escherichia coli composed of phosphatidylethanolamine, phosphatidylglycerol and cardiolipin. One type of vesicle contains trans-Δ9-octadecenoate, the other type trans-Δ9-hexadecenoate as predominant acyl chain component. The vesicles show order?disorder transitions at transition temperatures, Tt = 42° C and Tt = 29° C, respectively. A mixture of these vesicles is incubated at 45° C and lipid transfer is studied as a function of time using the phase transition as an indicator. The system reveals the following properties: Lipids are transferred between the two vesicle types giving rise to a vesicle population where both lipid components are homogeneously mixed. Lipid transfer is asymmetric, i.e. trans-Δ9-hexadecenoate-containing lipid molecules appear more rapidly in the trans-Δ9-octadecenoate-containing vesicles than vice versa. At a given molar ratio of the two types of vesicles the rate of lipid transfer is independent of the total vesicle concentration. It is concluded that lipid exchange through the water phase by way of single molecules or micelles is the mode of communication of these negatively charged lipid vesicles.  相似文献   

9.
(1) The polymorphic phase behaviour of aqueous dispersions of various synthetic phosphatidylethanolamines, both singly and in mixtures, has been investigated by 31P-NMR. (2) 14:014:0 PE remains in the lamellar phase up to 90°C. 18:1t18:1t PE exhibits a lamellar to hexagonal (HII) transition between 60°C and 63°C. For 18:1c18:1c PE, the lamellar to hexagonal (HII) transition occurs between 7 and 12°C, whereas for 18:2c18:2c PE, the hexagonal (HII) phase is the preferred structure above ?15°C. (3) Mixtures of 18:1c18:1c PE and 18:1t18:1t PE exhibit near-ideal miscibility behaviour. For mixtures of 18:1c18:1c PE and 14:014:0 PE there is evidence of fluid-solid immiscibility at temperatures below the gel-liquid crystalline transition temperature of the 14:014:0 PE component. Mixtures of 18:2c18:2c PE and 18:1t18:1t PE exhibit complex phase behaviour involving limited fluid-solid immiscibility at low temperatures and formation of a phase allowing isotropic motional averaging at higher temperatures. (4) 31P-NMR provides a graphic method for investigating the miscibility properties of mixed PE systems.  相似文献   

10.
Dipalmitoylphosphatidylcholine (DPPC) dispersed in perdeuterated glycerol was investigated in order to determine the effects on the Raman spectra of hydrocarbon chain interdigitation in gel-phase lipid bilayers. Interdigitated DPPC bilayers formed from glycerol dispersions in the gel phase showed a decrease in the peak height intensity I2850/I2880 ratio, for the symmetric and asymmetric methylene CH stretching modes, respectively, as compared to non-interdigitated DPPC/water gel-phase dispersions. The decrease in this spectral ratio is interpreted as an increase in chain-chain lateral interactions. Spectra recorded in the 700–740 cm?1 CN stretching mode region, the 1000–1200 cm?1 CC stretching mode region and the 1700–1800 cm? CO stretching mode region were identical for both the interdigitated and non-interdigitated hydrocarbon chain systems. At low temperatures the Raman peak height intensity ratios I2935/I2880 were identical for the DPPC/glycerol and DPPC/water dispersions, indicating that this specific index for monitoring bilayer behavior is insensitive to acyl chain interdigitation. The increase, however, in the change of this index at the gel-liquid crystalline phase transition temperature for the DPPC/glycerol dispersions implies a larger entropy of transition in comparison to the non-interdigitated DPPC/water bilayer system.  相似文献   

11.
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   

12.
Kinetic studies on the RNase T1-catalyzed transesterification of 12 dinucleoside monophosphates, Np1N2 (N1 = A, C, and U; N2 = A, C, G, and U) at pH 5, 25 °C, and 0.2 m ionic strength, revealed that the catalytic efficiency (kcatKm) for GpN substrates (H. L. Osterman, and F. G. Walz, Jr., 1978, Biochemistry, 17, 4142) was ~106-fold greater than corresponding ApNs and at least 108-fold greater than corresponding CpNs and UpNs. The catalytic activity with ApN substrates survives phenol extraction which indicates (along with other criteria) that it is intrinsic to RNase T1 and is not due to trace contamination by other nucleases. Circumstantial evidence is presented which suggests that homologous GpN and ApN substrates bind productively at different sites on the enzyme. The results of steady-state kinetic studies of RNase T1 with IpNs (N = C and U) were compared with those for GpNs and indicated that the primary effect of the guanine 2-NH2 group is to enhance substrate binding at the primary recognition site by ~2.6 kcal/mol. Values of (kcatKm) showed the order NpC > NpU (N = A, G, and I) which evidences the existence of a subsite for the leaving nucleoside group that prefers cytidine: interactions at this subsite are reflected in kcat rather than Km.  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.
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   

15.
16.
Systematic heat of dilution studies of the self-association of flavin mononucleotide (FMN) have been conducted as a function of ionic strength (0.05 – 2.0 m) and pH (5–9) in aqueous solution. The data are adequately described by the expression QT = ΔH ? (ΔHK)12 (QTcT)12 for an isodesmic self-association. QT is the molar heat of dilution, ΔH and K are the derived enthalpy and equilibrium constants for the process FMN + (FMN)i?1 ? (FMN)i, and cT is the concentration of FMN expressed in monomer units. Typical values derived for the various thermodynamic parameters at 25 °C are ΔG = ?3.56 kcal mol?1, ΔH = ?3.72 kcal mol?1, and ΔS = ?0.54 cal (mol · deg)?1. These data, plus nuclear magnetic resonance evidence (Yagi, K., Ohishi, N., Takai, A., Kawano, K., and Kyogoku, Y., 1976, Biochemistry15, 2877–2880) argue in favor of an open-ended association of flavin molecules. The signs of the various thermodynamic parameters suggest that both hydrophobic and surface energy forces contribute significantly to the association, while the lack of any significant ionic strength dependence indicates the lack of any ionic centers in the association.  相似文献   

17.
The Michaelis-Menten parameters, JM and Km of the initial 1-min fluxes of uptake of l-phenylalanine and of α-aminoisobutyric acid were determined for extracellular concentrations of Na+ ranging from 0.5 to 110 mequiv/l for Ehrlich ascites tumor cells. The maximal initial flux, JM, decreased with decrease in extracellular Na+ for both α-aminoisobutyric acid and phenylalanine but the Km for α-aminoisobutyric acid increased markedly as the Na+ concentration fell whereas the Km for phenylalanine decreased. Cycloleucine behaved like phenylalanine.The data provides strong evidence that the Na+-independent flux of phenylalanine is an exchange diffusion flux that can be varied by changing the intracellular level of amino acids such as phenylalanine. For phenylalanine, cyclolcucine, and methionine this exchange diffusion flux appears to be additive with the Na+-dependent initial flux. α-Aminoisobutyric acid also has an exchange diffusion that is Na+-independent but it has a high Km and is not additive with the Na+-dependent flux.  相似文献   

18.
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   

19.
The interactions between calmodulin, ATP and Ca2+ on the red cell Ca2+ pump have been studied in membranes stripped of native calmodulin or rebound with purified red cell calmodulin. Calmodulin stimulates the maximal rate of (Ca2+ + Mg2+)-ATPase by 5–10-fold and the rate of Ca2+-dependent phosphorylation by at least 10-fold. In calmodulin-bound membranes ATP activates (Ca2+ + Mg2+)-ATPase along a biphasic concentration curve (Km1 ≈ 1.4 μM, Km2 ≈ 330 μM), but in stripped membranes the curve is essentially hyperbolic (Km ≈ 7 μM). In calmodulin-bound membranes Ca2+ activates (Ca2+ + Mg2+)-ATPase at low concentrations (Km < 0.28 μM) in stripped membranes the apparent Ca2+ affinities are at least 10-fold lower.The results suggest that calmodulin (and perhaps ATP) affect a conformational equilibrium between E2 and E1 forms of the Ca2+ pump protein.  相似文献   

20.
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   

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