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.
α-Lactalbumin and dimyristoyl phosphatidylcholine were used as a prototype to study the influence of a protein conformational change, induced by the pH, on the interaction between that protein and a phospholipid.The enthalpy changes associated with the interaction of α-lactalbumin with dimyristoyl phosphatidylcholine vesicles were measured as a function of the molar ratio of phospholipid to protein, pH and temperature. Gel-filtration, electron-microscopic and fluorescence data for the same experimental conditions were also obtained. At pH 4 and 5, the enthalphy changes (ΔH) are not only larger than at physiological pH, but also show a maximum at about 23°C in the ΔH vs. temperature graph. At pH 6 and 7, on the contrary, ΔH increases with decreasing temperature without a maximum in the curve. Gel-chromatographic and electron-microscopic data show that at pH 6 and 7, the morphological characteristics of the vesicles are unchanged upon addition of α-lactalbumin, while at pH 4 and 5 at 23°C an extra peak appears in the gel-filtration graphs between the pure vesicles and α-lactalbumin. The new fraction contains lipid-protein complexes. Electron micrographs show that bar-shaped entities are formed. A red shift at 23°C and a blue shift at 37°C, both to 336 nm, are observed for λmax of the fluorescence emission spectra at pH 4 when α-lactalbumin is brought into contact with the phospholipid. At the same time, a strong increase in the fluorescence intensity is observed. The chromatographic and fluorescence data indicate that a lipid-protein complex with a molar ratio of approx. 80 is formed. At pH 7 and different temperatures, the emission maximum remains at the wavelength of pure α-lactalbumin, the change in the fluorescence intensity, however, indicates that interaction with the lipid occurs.The results can be explained on the basis of an electrostatic interaction at pH 6 and 7, and a hydrophobic interaction at pH 4 and 5.  相似文献   

2.
Ventilation was measured directly in the hagfish, Myxine glutinosa L., by means of an electro-magnetic blood flowmeter. Ventilatory flow and frequency increased from 0.86 ± 0.27 ml·min?, and 18.2 ± 5.1·min?, respectively, at 7°C to 1.70 ± 0.20 ml·min?, and 70.1 ± 9.5·min? at 15 ·C.Standard oxygen consumption,V?O2, was measured in non-buried hagfish. V?O2 was 0.57 ± 0.17μl O2·g?1·min?1 at 7°C, and 0.85 ± 0.12μl O2·g?1·min?1 at 15°C.  相似文献   

3.
Treatment of the S3G strain of HeLa cells with dexamethasone inhibits cholesterol synthesis and thus results in decreased plasma membrane cholesterol-to-protein ratios. Incubation of HeLa cells with dexamethasone for 24 h lowers the steady-state fluorescence polarization of 1,6-diphenyl-1,3,5-hexatriene (DPH) in intact cell plasma membranes and isolated plasma membranes (Johnston, D. and Melnykovych, G. (1980) Biochim. Biophys. Acta 596, 320–324). We have examined the effect of dexamethasone treatment of S3G HeLa cells on the lateral diffusion of the fluorescent lipid analogue 3,3′-dioctadecylindocarbocyanine iodide (DiI) by the fluorescence photobleaching recovery technique. The lateral diffusion of DiI was measured in cells 0, 2, 6, 12, and 24 h following treatment with dexamethasone and in cells identically handled without dexamethasone at 37°C. The diffusion constants of DiI in the treated and untreated cell membranes at zero time were (4.52±0.30) · 10?9 cm2/s and (4.56±0.24) · 10?9 cm2/s, respectively. There was no significant change in the lateral diffusion of DiI in the untreated cells over the 24 h period. The lateral diffusion of the lipid probe in the dexamethasone-treated cells began to increase 6 h following treatment and reached (6.43±0.27) · 10?9 cm2/s at 24 h. The lateral diffusion of DiI was also measured at 25, 17, 10 and 4°C following 24 h incubation with and without dexamethasone. The effect of dexamethasone treatment on the lipid probe lateral diffusion observed at 37°C is decreased at 25°C and reversed in direction at 10 and 4°C. These results agree with those obtained in artificial systems containing varying amounts of cholesterol and support the suggestion that cholesterol acts to suppress phospholipid phase changes in animal cells. The lateral diffusion of DiI localized as a monolayer at a mineral oil-water interface was measured by fluorescence photobleaching recovery. The resulting data and the viscosity of the mineral oil were used to calculate the microviscosities of the plasma membranes of untreated and dexamethasone-treated cells at 25°C. Membrane microviscosities were also calculated from the fluorescence polarization studies cited above. In both cases the dexamethasone treatment reduced the apparent microviscosity by approximately 25%. However, the absolute microviscosity values obtained by the two techniques differ by a factor of 3.  相似文献   

4.
We investigated the interaction between α-lactalbumin and sonicated dimyristoylphosphatidylcholine at pH 4 and different temperatures. (1) At 23°C and lipid-to-protein molar ratios below 170, the interaction results in a disruption of the original vesicles to form smaller complex particles. By the sedimentation velocity method we determined for this particle a molar mass of (1.05 ± 0.16) · 106 g·mol?1. The lipid-to-protein molar ratio within the complex particle is 70/1, as earlier estimated. It follows that there are approximately 1200 lipid and 17 α-lactalbumin molecules per particle. At molar ratios above 170, α-lactalbumin strongly associates with the vesicles. In this case the vesicle entity remains. The ability of α-lactalbumin to break up the vesicles at this temperature is determined by the number of protein molecules which are required in the complex particle. (2) By means of fluorescence polarization of the lipophilic probe 1,6-diphenyl-1,3,5-hexatriene and energy transfer of the tryptophan groups of the protein to 1,3-(1,1′-dipyrenyl)propane located in the hydrocarbon region of the vesicles, it is shown that with increasing temperature above 25°C, complexes of decreasing internal lipid-to-protein molar ratio are formed. However, by electron microscopy we show that the overall size of these complexes remains approximately the same, i.e., bars with dimensions 70 × 220 A?. A temperature-reversible transformation occurs between these complexes, which cannot be isolated by gel chromatography. In contrast, the complex of molar ratio 70/1 remains stable at lower temperatures.  相似文献   

5.
Glucagon can interact rapidly with multilamellar vesicles of dimyristoyl glycerophosphocholine over a narrow temperature range around or above the phase transition temperature of the pure phospholipid. The temperature dependence of the rates arises, in large part, from glucagon-induced alterations in the phase transition properties of the phospholipid. Similar effects are observed with dilaury glycerophosphocholine but the rate of reaction of glucagon with multilamellar dipalmitoyl glycerophosphocholine is too slow to measure.The rate of reaction of glucagon with equimolar mixtures of two phospholipid molecules has also been studied. Mixtures of dilauryl glycerophosphocholine and distearoyl glycerophosphocholine are known to exhibit lateral phase separation in the gel state. The presence of distearoyl glycerophosphocholine has no effect on the rate of reaction with glucagon, despite the increased number of phase boundaries present. In the case of mixtures of dilauryl glycerophosphocholine and dimyristoyl glycerophosphocholine, glucagon appears to induce some lateral phase separation. This is demonstrated by the ability of glucagon to react rapidly with this lipid mixture, even at temperatures well below the phase transition temperature of the mixture and by differential scanning calorimetry.The thermodynamics of the binding of glucagon to dimyristoyl glycerophosphocholine and dilauryl glycerophosphocholine were analyzed with Scatchard plots calculated from measurements of the fluorescence enhancement caused by lipids. Equilibrium binding constants of glucagon to dimyristoyl glycerophosphocholine and dilauryl glycerophosphocholine are 1·105 and 5·104 M?1, respectively. These values are relatively insensitive to temperature, indicating that the equilibrium being measured is between lipid-bound glucagon and free lipid which has had its phase transition properties altered. The number of moles of lipid bound per mole of glucagon decreases markedly above the phase transition temperature. In the water-soluble complex formed between glucagon and dimyristoyl glycerophosphocholine, the peptide binds directly to only 40% of the lipid molecules but, nevertheless, is able to modify the phase transition properties of all of the lipid in the particle.  相似文献   

6.
Glycophorin from human erythrocytes has been incorporated into liposomes of dimyristoylphosphatidylcholine (DMPC), dipalmitoylphosphatidylcholine (DPPC) and distearoylphosphatidylcholine (DSPC). The thermal properties of unsonicated liposomes with glycophorin/lipid molar ratios up to 4·10?3 have been studied by differential scanning calorimetry and the numbers of lipids withdrawn from participation in the gel-to-lamellar phase transition were found to be 42±22 (DMPC), 197±28 (DPPC) and 240±64 (DSPC). The initial rates of agglutination of sonicated liposomes with glycophorin/lipid molar ratios up to 4·10?3 by wheat germ agglutinin in the concentration range 0–7 μM have been measured over a range of temperature. Below the gel-to-lamellar phase transition (Tc) the rates of agglutination increase with acyl chain length in the sequence DMPC < DPPC < DSPC. Agglutination is found to be second order in liposome concentration and is completely reversed on saturation of the wheat germ agglutinin-binding sites by N-acetylglucosamine. Agglutination rates decrease with increasing temperature below Tc and are largely independent of temperature above Tc. The results are discussed in relation to the clustering of glycophorin in the phospholipid bilayers and its effect on binding and subsequent interliposomal bridge formation by wheat germ agglutinin.  相似文献   

7.
The isolation and characterization of a hybridoma cell line producing a monoclonal IgG1 antibody against a spin-label nitroxide group is described. The antibody recognizes a synthetic hapten containing linked dinitrophenyl and 2,2,6,6-tetramethylpiperidinyl 1-oxy groups, having an affinity of 3.6±1.0·106 M?1 for the soluble hapten at 25°C. The antibody binds to phospholipid vesicles containing 2 mol% of spin label-derivitized lipid (lipid hapten) with an affinity of 1.5±0.2·108 M?1. This monoclonal IgG1 mediates the binding of hapten-bearing lipid vesicles to mouse macrophage RAW264 cells bearing Fc receptors. The cellular responses to this binding are similar to those observed previously using polyclonal rabbit anti-hapten IgG. As with the heterogeneous antibodies, the monoclonal IgG1 is more efficient in mediating cellular uptake when the vesicles are in the ‘fluid’ physical state (dimyristoylphosphatidylcholine at 37°C) compared to ‘solid’ (dipalmitoylphosphatidylcholine at 37°C). Despite the enhanced binding of ‘fluid’ phospholipid vesicles to cells, only the ‘solid’ vesicles triggered a significant respiratory burst in RAW264 macrophages.  相似文献   

8.
The kinetics of the chlorophyll fluorescence rise induced by adding 20 mM MgCl2 to a suspension of isolated pea chloroplasts treated with 3-(3,4-dichlorophenyl)-1, 1-dimethylurea (DCMU) have been examined experimentally and theoretically as a function of temperature. The application of similarity arguments and particle aggregation theory to the experimental results suggests that at the first approximation, the salt-induced time-dependent fluorescence changes may be described by the diffusion-controlled lateral movement of Photosystem II pigment-protein complexes. From an analysis of the temperature dependence of the fluorescence changes, estimates obtained for the lateral diffusion coefficients were 1.85 · 10?12–3.08 · 10?11 cm2/s over the temperature range 10°C ? T?30°C.  相似文献   

9.
Measurements of the equilibrium and temperature-jump u.v., visible, and induced c.d. spectra of Methyl Orange (MO) in the presence of cyclomalto-octaose (γ-cyclodextrin, γ-CD) have been carried out. Three mechanistic steps were detected through the temperature-jump data (25.0°):
where K1, K2, and K3 are 45 (±7), 2.0 (±1.1) × 106, and 6.1 (±2.5) × 103 dm3.mol?1, respectively, k2 = 9.4 (±5.1) × 109 dm3.mol?1.s?1, and k?2 = 4.8 (±0.8) × 103 s?1. The equilibrium u.v./visible data are also consistent with this reaction scheme. The high stability of the dimer inclusion complex (MO)2 · γ-CD compared to that of the monomer inclusion complex MO · γ-CD appears to be related to the annular diameter of γ-CD and demonstrates a degree of selectivity in cyclodextrin inclusion complexes. The (MO)2 · (γ-CD)2 complex also contains a dimer, included by both γ-CD molecules.  相似文献   

10.
In all larval stages of Carcinus maenas L. oxygen consumption was measured at three temperatures (12,18,25 °C). Values increased during development and were in the range of 0.037 ± 0.01 (zoea-1, 12°C, x? ± 95% CL) to 0.734 ± 0.047 μl O2 · h?1 · ind?1 (megalopa, 25 °C). Growing larvae showed temperature dependent trends in weight specific respiration rates (referred to dry wt; DW), with values between ≈2.4 and 9.4 μl O2· h?1·mg DW?1. Increase in oxygen consumption of megalops did not differ much at temperatures between 18 and 25 °C. This points to an exceptional physiological position of this stage. Fed zoea-1 of C. maenas (18 °C) revealed growth rates in terms of 40% DW, 20% carbon (C), 30% nitrogen (N) and 65% hydrogen (H). At the same time larvae gained individual energy by 13% (J · ind?1), while weight specific energy dropped by ≈ 19% (J · mg DW?1) during the first day and remained constant until the moult. Starved zoea-1 of C. maenas (18 ° C) gained ≈ 20 % in DW through the first day, probably caused by inorganic salts which enter the organism after the moult of the prezoea. DW dropped to ≈ 25 % of initial value, when starvation continued. Single components decreased by ≈50% (C), 54% (N), 57% (J · ind?1). Weight specific energy (J · mg DW?1) decreased by 40% during the first 4 days of starvation, remaining constant thereafter. Individual respiration rate (R) dropped by 61 %, weight specific respiration rate (QO2) by 55 %. Individual energy loss in starved zoea-1 was 0.077 J over a period of 11 days. In this period ≈ 9.3 μl O2·ind?1 were consumed. Thus effective oxygen capacity was lower than in growing larvae. It dropped to 5.3 J·mlO2?1 after 4 days and remained constant if starvation continued, i.e. 65 % of possible energy loss occurred during the first 4 days. Decrease in requirement for oxygen and its effective capacity were both recognized as independent components of survival during starvation. Partitioning of energy through individual larval development of C. maenas was investigated for all five larval stages. The cumulative budget could be calculated: consumption (C) = 28.23 J, growth (G) = 0.92 J, exoskeleton (Ex) = 0.20 J, metabolism (M) = 5.30 J, egestion and excretion (E) = 21.82 J. Mean gross and net growth efficiency were, K1 = 3.3% and K2 = 14.8%, respectively.  相似文献   

11.
We have investigated the binding of 2′-deoxyuridine 5′-monophosphate (2′-dUMP) to Leishmania major deoxyuridine 5′-triphosphate nucleotide hydrolase (dUTPase) by isothermal titration microcalorimetry under different experimental conditions. Binding to dimeric L. major dUTPase is a non-cooperative process, with a stoichiometry of 1 molecule of 2′-dUMP per subunit. The utilization of buffers with different ionization enthalpies has allowed us to conclude that the formation of the 2′-dUMP–dUTPase complex, at pH 7.5 and 30 °C, is accompanied by the uptake of 0.33±0.05 protons per dUTPase subunit from the buffer media. Moreover, 2′-dUMP shows a moderate affinity for the enzyme, and binding is enthalpically driven across the temperature range studied. Besides, whereas ΔG° remains practically invariant as a function of temperature, both ΔH and ΔS° decrease with increasing temperature. The TS and TH were 23.4 and 13.6 °C, respectively. The temperature dependence of the enthalpy change yields a heat capacity change of ΔCp°=?618.1±126.4 cal·mol?1·K?1, a value low enough to discard major conformational changes, in agreement with the fitting model. An interpretation of this value in terms of solvent-accessible surface areas is provided.  相似文献   

12.
V.A. Sineshchekov  F.F. Litvin 《BBA》1977,462(2):450-466
Red luminescence of purple membranes from Halobacterium halobium cells in suspension, dry film or freeze-dried preparations was studied and its emission, excitation and polarization spectra are reported. The emission spectra have three bands at 665–670, 720–730 and at 780–790 nm. The position (maximum at 580 nm) and shape of the excitation spectra are close to those of the absorption spectra. The spectra depend on experimental conditions, in particular on pH of the medium. Acidification increases the long wavelength part of the emission spectra and shifts the main excitation maximum 50–60 nm to the longer wavelength side. Low-temperature light-induced changes of the absorption, emission and excitation spectra are presented. Several absorbing and emitting species of bacteriorhodopsin are responsible for the observed spectral changes. The bacteriorhodopsin photoconversion rate constant was estimated to be about 1 · 1011 s?1 at ? 196°C from the quantum yields of the luminescence (1 · 10?3) and photoreaction (1 · 10?1). The temperature dependence of the luminescence quantum yield points to the existence of two or three quenching processes with different activation energies. High degree of luminescence polarization (about 45–47%) throughout the absorption and fluorescence spectra and its temperature independence show that there is no energy transfer between bacteriorhodopsin molecules and no chromophore rotation during the excitation lifetime. In carotenoid-containing membranes, energy migration from the bulk of carotenoids to bacteriorhodopsin was not found either. Bacteriorhodopsin phosphorescence was not observed in the 500–1100 nm region and the emission is believed to be fluorescence by nature.  相似文献   

13.
High-resolution NMR is shown to be applicable for investigation of membrane proteins and membrane-active peptides embedded into lipid-protein nanodiscs (LPNs). 15N-Labeled K+-channel from Streptomyces lividans (KcsA) and the antibiotic antiamoebin I from Emericellopsis minima (Aam-I) were embedded in LPNs of different lipid composition. Formation of stable complexes undergoing isotropic motion in solution was confirmed by size-exclusion chromatography and 31P-NMR spectroscopy. The 2D 1H-15N-correlation spectra were recorded for KcsA in the complex with LPN containing DMPC and for Aam-I in LPNs based on DOPG, DLPC, DMPC, and POPC. The spectra recorded were compared with those in detergent-containing micelles and small bicelles commonly used in high-resolution NMR spectroscopy of membrane proteins. The spectra recorded in LPN environments demonstrated similar signal dispersion but significantly increased 1HN line width. The spectra of Aam-I embedded in LPNs containing phosphatidylcholine showed significant selective line broadening, thus suggesting exchange process(es) between several membrane-bound states of the peptide. 15N relaxation rates were measured to obtain the effective rotational correlation time of the Aam-I molecule. The obtained value (~40 nsec at 45°C) is indicative of additional peptide motions within the Aam-I/LPN complex.  相似文献   

14.
A unique 14 base oligodeoxynucleotide dTTCATCAGCCACTG complementary to glucagon mRNA was deduced from the amino acid sequence of the hormone (residues 24–28; GLN-TRP-LEU-MET-ASN). The oligonucleotide specifically hybridized to RNA from rabbit pancreas and human pancreatic islet cells. No detectable hybridization was observed with either yeast or rat liver RNA. The melting temperature of the hybrids was 50 ± 5°C indicating no significant mismatch for human or rabbit glucagon mRNA. Hybridization kinetics followed a single pseudofirst-order reaction (Cot0 · 5 = 2.5 × 10?4 M sec). From the extent of reaction at completion there are a minimum of 43 fmol glucagon mRNA/mg RNA (total pancreas).  相似文献   

15.
Sonicated dispersions of 1,2-dipalmitoylsn-glycero-3-phosphorylcholine and of 1,3-dipalmitoylglycero-2-phosphorylcholine were examined by proton nuclear magnetic resonance (NMR) as a function of temperature. The —(CH2)n)— peak in the spectrum of the sn-3-isomer of dipalmitoylphosphatidylcholine showed the characteristic dramatic changes in the peak intensity and width associated with the phase transition between the liquid crystalline and gel states of the phospholipid. This occurred over a 2–3°C temperature range with the midpoint of the transition at 38.5°C. With the 2-isomer the change in phase took place over a similar temperature range but the midpoint was at 33.8°C. This lower phase transition temperature is presumably the result of increased acyl chain mobility caused by the increased separation of the two acyl chains by the centre carbon of the glycerol backbone. The effect of sonication of the broadening of the range and lowering of the midpoint temperature of the phase transition from that of the corresponding unsonicated dispersions was similar with each isomer. This suggests that the overall geometry of the sonicated vesicles of the isomers is similar.  相似文献   

16.
Mark A. Jensen  Philip J. Elving 《BBA》1984,764(3):310-315
The rate constant, kd, for the dimerization of the free radical (NAD·), produced on the initial one-electron reduction of NAD+, was measured by double potential-step chronoamperometry, fast-scan cyclic voltammetry (cathodic-anodic peak current ratio) and slow-scan cyclic voltammetry (peak potential shift) for a medium in which neither NAD+ nor its reduction products are adsorbed at the solution/electrode interface. All three methods give concordant values of kd (approx. 3·107 M?1·s?1), which are in reasonable accord with the values determined by pulse radiolysis but are considerably greater than values previously determined electrochemically. For the NAD+/NAD· couple, the heterogeneous rate constant (ks,h) exceeds 1 cm·s?1 at 25°C and the formal potential (E0c) vs. sce is ? 1.155 V at 25°C and ? 1.149 V at 1°C at pH 9.1, with an uncertainty of about ±0.005 V.  相似文献   

17.
Telemetered heart rate (fH) was examined as an indicator of activity and oxygen consumption rate (VO2) in adult, cultivated, Atlantic salmon, Salmo salar L. Heart rate was measured during sustained swimming in a flume for six fish at 10° C [mean weight, 1114 g; mean fork length (f. l.), 50·6 cm] and seven fish at 15° C (mean weight, 1119 g; mean f. l., 50·7 cm) at speeds of up to 2·2 body lengths/s. Semi–logarithmic relationships between heart rate and swimming speed were obtained at both temperatures. Spontaneously swimming fish in still water exhibited characteristic heart rate increases associated with activity. Heart rate and Vo2 were monitored simultaneously in a 575–1 circular respirometer for six fish (three male, three female) at 4° C (mean weight, 1804 g; mean F. L., 62· cm) and six fish (three male, three female) at 10° C (mean weight, 2045 g; mean f. l., 63·2 cm) during spontaneous but unquantified activity. Linear regressions were obtained by transforming data for both fH and Vo2 to log values. At each temperature, slopes of the regressions between fH and Vo2 for individual fishes were not significantly different, but in some cases elevations were. All differences in elevation were between male and female fish. There were no significant differences in regression slope or elevation for fish of the same sex at the two temperatures and so regressions were calculated for the sexes, pooling data from 4 and 10° C. There was no significant difference in the mean ± S. D. Vo2 between the sexes at 4° C (male, 66·0 ± 59·6 mgO2 kg?1 h?1; female, 88·0 ± 60·1 mgO2 kg?1 h?1) or 10° C (male, 166·2 ± 115·4 mgO2 kg?1 h?1; female, 169·2 ± 111–1 mgO2 kg?1h?1). Resting Vo2 (x?± s. d.) at 4°C was 36·7 ± 8.4 mgO2 kg?1 h?1, and 10° C was 72·8 ± 11·9 mgO2 kg?1 h?1. Maximum Vo2 (x?± S. D.) at 4° C was 250·6 ± 40·2 mgO2 kg?1 h?1, and at 10° C was 423·6 ± 25·2 mgO2 kg?1 h?1. Heart rate appears to be a useful indicator of metabolic rate over the temperature range examined, for the cultivated fish studied, but it is possible that the relationship for wild fish may differ.  相似文献   

18.
The binding of methyl isonitrile (CH3Nandz.tbnd;C) to hemoglobin β chains has been studied by measuring the 1H nuclear magnetic resonance transverse relaxation times for methyl isonitrile as a function of protein concentration, temperature and 14N decoupling. Binding of methyl isonitrile both at the heme iron and at a non-specific site (or sites) has an effect upon the measured nuclear spin relaxation times. The results yield a value of 57 ± 12 seconds?1 (20 °C) for the “off” rate constant K?1 for specific binding and an Arrhenius activation energy for k?1 of 14 ± 3 kcal mol?1.  相似文献   

19.
Melting and crystallization phenomena of amylose-lipid complexes, crystallized either from dilute solution or concentrated amylose ‘melts’ under various conditions, were studied using differential scanning calorimetry (DSC). The melting enthalpies (complex/H2O: 0·1–0·5 w/w) of the solution-grown crystalline complexes were 20·4±0·8 J g?1 for amylosemonopalmitin (AM-1-C16), 26·5±1·5 J g?1 for amylose-lysolecithin (AM-lys/in) and 26·6±1·6 J g?1 for amylose-lauric acid (AM-C12). While melting of the AM-lys/in complex showed a single transition for all concentrations studied, the melting behavior of the AM-1-C16 and the AM-C12 was rather complex at low or intermediate water contents. At a heating rate of 10°C min?1 two endothermic transitions with an intermediate exothermic peak were observed, indicative of non-equilibrium melting. A process of partial melting, followed by recrystallization and final melting, is suggested to explain such multiple-melting characteristics. These phenomena become less prominent with increasing water content; presumably due to the depression of the glass transition (Tg) and the melting temperatures (Tm). The annealing behavior of AM-1-C16 further suggested that the development of new structural order upon heating takes place primarily after partial melting of the initial crystalline structure. Overall, the DSC data are typical of those for semicrystalline polymers and consistent with a lamellar type of molecular organization in the crystalline regions of these materials.  相似文献   

20.
A methodological study has been made with a syringe titration unit attached to an LKB batch microcalorimeter. The presicion and accuracy of the instrument assembly have been evaluated by neutralization reactions and by dilution of sucrose solutions. As an example, heat quantities on the order of 10 mJ accompanying the addition of 10 μl titrant solution could be determined with an accuracy of better than 1%. A stepwise titration procedure was used to characterize the binding of indole-3-propionic acid to α-chymotrypsin. The following thermodynamic data were obtained (25°C, acetate buffer, pH 5.80): ΔG0 = ?18.46±0.17 kJ·mol?1, ΔH0 = ?15.26±0.20 kJ·mol?1, ΔS0 = 10.85±1.21 JK?·mol?1.  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