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 29 毫秒
1.
The phases and transition sequences for aqueous dispersions of mixtures of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) and 1,2-dipalmitoyl-sn-glycerol (1,2-DPG) have been studied by differential scanning calorimetry, dynamic x-ray diffraction, freeze-fracture electron microscopy, 31P-nuclear magnetic resonance spectroscopy, and Fourier-transform infrared spectroscopy. The results have been used to construct a dynamic phase diagram of the binary mixture as a function of temperature over the range 20 degrees-90 degrees C. It is concluded that DPPC and 1,2-DPG form two complexes in the gel phase, the first one with a DPPC/1,2-DPG molar ratio of 55:45 and the second one at a molar ratio of approximately 1:2, defining three different regions in the phase diagram. Two eutectic points are postulated to occur: one at a very low 1,2-DPG concentration and the other at a 1,2-DPG concentration slightly higher than 66 mol%. At temperatures higher than the transition temperature, lamellar phases were predominant at low 1,2-DPG concentrations, but nonlamellar phases were found to be predominant at high proportions of 1,2-DPG. A very important aspect of these DPPC/1,2-DPG mixtures was that, in the gel phase, they showed a ripple structure, as seen by freeze-fracture electron microscopy and consistent with the high lamellar repeat spacings seen by x-ray diffraction. Ripple phase characteristics were also found in the fluid lamellar phases occurring at concentrations up to 35.6 mol% of 1,2-DPG. Evidence was obtained by Fourier transform infrared spectroscopy of the dehydration of the lipid-water interface induced by the presence of 1,2-DPG. The biological significance of the presence of diacylglycerol in membrane lipid domains is discussed.  相似文献   

2.
The thermodynamics of the lipase-catalyzed esterification of glycerol with n-octanoic acid have been investigated with acetonitrile, benzene, and toluene as solvents and in the neat reaction mixture (no organic solvent added). This esterification reaction leads to five products: 1-monooctanoyl glycerol, 2-monooctanoyl glycerol, 1,2-dioctanoyl glycerol, 1,3-dioctanoyl glycerol and 1,2,3-trioctanoyl glycerol. This, in turn leads to a total of 12 reactions. Values of the equilibrium constants for these reactions have been measured (HPLC, GC, and LC/MS) at 37°C in the above mentioned media. The equilibrium constants range from 0.9 to 20.7, 0.20 to 8.0, 0.23 to 10.0, and 0.57 to 2.2 in acetonitrile, benzene, toluene, and neat media, respectively. Relative standard molar Gibbs free energies of formation ΔfGm0 of 1-monooctanoyl glycerol, 2-monooctanoyl glycerol, 1,2-dioctanoyl glycerol, 1,3-dioctanoyl glycerol and 1,2,3-trioctanoyl glycerol in the organic solvents and in the neat reaction mixture have been calculated and used to compactly summarize the thermodynamics of these reactions. The results show an approximate correlation with the permittivities of the solvents.  相似文献   

3.
Cytochalasin B (CB) was prepared by methanol extraction of dehydrated mold (Drechslera dematioidea) matte, reverse-phase C18 silica gel batch adsorption, selective elution with 1:1 (v/v) hexane:tetrahydrofuran (THF), crystallization, preparative TLC, and recrystallization. Unit gravity silica gel normal phase chromatography afforded additional CB. Yield per liter of medium was 300 mg of CB greater than 95% pure by NMR, HPLC (60:40 hexane:THF, Lichrosorb Si60 silica gel, 230 nm), and TLC. CB added exogenously to mouse organs at 1 and 5 micrograms/organ was recovered 70 to 100% by methanol extraction, adsorption to C18 silica gel Sep-Pak cartridges, elution with ethyl acetate, and analysis by TLC and/or HPLC. Limiting sensitivity (micrograms/extract) was 0.5 TLC; 1.0 HPLC. Quantitative extraction was confirmed with 3H-labeled CB. CB ip in mice at 50 mg/kg (LD10) distributed rapidly into liver, renal fat, kidney, intestines, mesentery, pancreas, spleen, and blood cells and was cleared from all but liver within 24 h. CB was below detectable levels in thymus, lymph nodes, heart, brain, bone marrow, and lungs. Cytochalasin A is fixed to tissues and not extractable. This work affords a source of CB in quantities permitting in vivo study, provides methods for extraction and analysis, and reveals the pharmacokinetics of ip bolus CB.  相似文献   

4.
The kinetics of 2,3-diphosphoglycerate (2,3-DPG) net breakdown was examined in intact human erythrocytes incubated at pH 7.00 and 37 °C. The concentrations of 2,3-DPG, 1,3-diphosphoglycerate (1,3-DPG), 3-phosphoglycerate, ATP, Pi, glucose, and lactate were determined during 10 to 12 h. Since the concentration of 1,3-DPG has been suggested to be the main regulating factor with respect to the rate of 2,3-DPG net breakdown the interdependence between the concentration of 1,3-DPG and pH was determined in the range of pH 6.9 to 7.4. It was found that the stationary level of 1,3-DPG decreased strongly with decreasing pH within this range. Qualitatively, the net breakdown of 2,3-DPG observed at pH 7.00 can be explained by the lowered level of 1,3-DPG. The influence of the concentration of Pi upon the rate of net degradation of 2,3-DPG at pH 7.00 was studied at low cell volume fraction (0.04), where given concentrations of Pi could be maintained for several hours. A marked increase in the rate of 2,3-DPG net breakdown by Pi was demonstrated. Computer simulations showed that activation of diphosphoglycerate phosphatase by the increasing concentration of Pi and decrease of degree of inhibition of the diphosphoglycerate mutase by the decreasing concentration of 2,3-DPG may well keep the rate of the degradation balanced at the time constant value observed. On the basis of the observed kinetics and a computer simulation, the flux through the phosphoglycerate bypass was estimated to be 10 to 15% of the total glycolytic flux at physiological conditions.  相似文献   

5.
To obtain predominant bacteria degrading crude oil, we isolated some bacteria from waste soybean oil. Isolated bacterial strain had a marked tributyrin (C4:0) degrading activity as developed clear zone around the colony after incubation for 24h at 37 degrees C. It was identified as Klebsiella sp. Y6-1 by analysis of 16S rRNA gene. Crude biosurfactant was extracted from the culture supernatant of Klebsiella sp. Y6-1 by organic solvent (methanol:chloroform:1-butanol) after vacuum freeze drying and the extracted biosurfactant was purified by silica gel column chromatography. When the purified biosurfactant dropped, it formed degrading zone on crude oil plate. When a constituent element of the purified biosurfactant was analyzed by TLC and SDS-PAGE, it was composed of peptides and lipid. The emulsification activity and stability of biosurfactant was measured by using hydrocarbons and crude oil. The emulsification activity and stability of the biosurfactant showed better than the chemically synthesized surfactant. It reduced the surface tension of water from 72 to 32 mN/m at a concentration of 40 mg/l.  相似文献   

6.
2,3-Diphosphoglycerate (2,3-DPG) has been found to be the product of the enzymatic degradation of cyclic 2,3-diphosphoglycerate (cDPG) in the archaebacterium Methanobacterium thermoautotrophicum delta H. Although 2,3-DPG has not previously been detected as a major soluble component of M. thermoautotrophicum, large pools accumulated at an incubation temperature of 50 degrees C (below the optimum growth temperature of 62 degrees C). Under these conditions, cellular activity was significantly decreased; a return of the culture to the optimum growth temperature restored the 2,3-DPG pool back to original low levels and caused steady-state cDPG levels to increase again. While 13CO2-pulse/12CO2-chase experiments at 50 degrees C showed that the cDPG turned over, the appearance of 2,3-DPG at NMR-visible concentrations required at least 10 h. Production of 2,3-DPG in vivo was prevented by exposure of the cells to O2. The enzyme responsible for this hydrolysis of cDPG was purified by affinity chromatography and appears to be a 33-kDa protein. Activity was detected in the presence of oxygen and was enhanced by a solution of 1 M KCl, 25 mM MgCl2, and dithiothreitol. Both Km and Vmax have been determined at 37 degrees C; kinetics also indicate that in vitro the product, 2,3-DPG, is an inhibitor of cDPG hydrolysis. These findings are discussed in view of a proposed role for cDPG in methanogens.  相似文献   

7.
Lipid activation of protein kinase C alpha (PKC alpha) was studied using a model mixture containing POPC/POPS (molar ratio 4:1) and different proportions of either DPG or POG. The lipid mixtures containing DPG were physically characterized by using different physical techniques, and a phase diagram was constructed by keeping a constant POPC/POPS molar ratio of 4:1 and changing the concentration of 1,2-DPG. The phase diagram displayed three regions delimited by two compounds: compound 1 (CO(1)) with 35 mol % of 1,2-DPG and compound 2 (CO(2)) with 65 mol % of 1,2-DPG. PKC alpha activity was assayed at increasing concentrations of 1,2-DPG, maximum activity being reached at 30 mol % 1,2-DPG, which decreased at higher concentrations. Maximum activity occurred, then, at concentrations of 1,2-DPG which corresponded to the transition from region 1 to region 2 of the phase diagram. It was interesting that this protein was maximally bound to the membrane at all DPG concentrations. Similar results were observed when the enzyme was activated by POG, when a maximum was reached at about 10 mol %. This remained practically constant up to 50 mol %, about which it decreased, the binding level remaining maximal and constant at all POG concentrations. The fact that in the assay conditions used maximal binding was already reached even in the absence of diacylglycerol was attributed to the interaction of the C2 domain with the POPS present in the membrane through the Ca(2+) ions also present. To confirm this, the isolated C2 domain was used, and it was also found to be maximally bound at all DPG concentrations and even in its absence. Since the intriguing interaction patterns observed seemed to be due then to the C1 domain, the PKC alpha mutant D246/248N was used. This mutant has a decreased Ca(2+)-binding capacity through the C2 domain and was not activated nor bound to membranes by increasing concentrations of DPG. However, POG was able to activate the mutant, which showed a similar dependence on POG concentration with respect to activity and binding to membranes. These data underline the importance of unsaturation in one of the fatty acyl chains of the diacylglycerol.  相似文献   

8.
The oligonucleotide 5'-d(TCTACGCGTTCT) reacts with trans-diamminedichloroplatinum(II) to yield primarily trans-[Pt(NH3)2[d(TCTACGCGTTCT)-N7-G(6),N7-G(8)]], containing the desired trans-[Pt(NH3)2[d(GCG)]] 1,3-cross-link. A key element of the platination reaction is the use of low pH to suppress coordination at A(4). The product was fully characterized by pH-dependent NMR titrations, enzymatic degradation analysis, and 195Pt NMR spectroscopy. Interestingly, the 1,3-cross-linked adduct is unstable at neutral pH, rearranging unexpectedly to form the linkage isomer trans-[Pt(NH3)2[d-(TCTACGCGTTCT)-N3-C(5),N7-G(8)]]. This rearrangement product is more stable than the initially formed isomer and could be characterized by pH-dependent NMR titrations, enzymatic degradation analysis, liquid secondary ion mass spectrometric analysis of an enzymatically digested fragment, 195Pt NMR spectroscopy, and modified Maxam-Gilbert footprinting experiments. By contrast, the 1,3-intrastrand cross-linked isomer rearranges during the course of both pH titration and enzymatic degradation experiments to form the 1,4-adduct. The equilibrium constant for this rearrangement is approximately 3, favoring the 1,4-adduct. Kinetic studies of the linkage isomerization reaction reveal t1/2 values for the first-order disappearance of the 1,3-intrastrand cross-linked isomer ranging from 129 (at 30 degrees C) to 3.6 h (at 62 degrees C), with activation parameters delta H not equal to = 91 +/- 2 kJ/mol and delta S not equal to = -58 +/- 8 J/(mol.K). Mechanistic implications of these kinetic results as well as the general relevance of this linkage isomerization reaction to platinum-DNA chemistry are briefly discussed.  相似文献   

9.
The interaction of diacylglycerols, primarily 1,2-dilauroyl-sn-glycerol (1,2-DLG), with egg phosphatidylcholine (PC) bilayers was studied by NMR spectroscopy and other physical techniques. In the low proportions used (less than or equal to 20 mol % with respect to total lipid), 1,2-DLG formed bilayers with PC with no hexagonal phase separation, as assessed by light, polarizing and electron microscopy, and 31P and 13C NMR spectroscopy. The 13C-carbonyl chemical shift of 90% [13C]carbonyl 1,2-DLG was monitored in small unilamellar vesicles as a function of relative DLG content (1.5-20%) and temperature (10-55 degrees C). The chemically inequivalent sn-1 and sn-2 carbonyls gave a single, narrow resonance in vesicles, in contrast to neat 1,2-DLG and 1,2-DLG in organic solvents, whose spectra showed two well-separated carbonyl resonances. The chemical shift of 1,2-DLG in PC shows that the carbonyl groups are proximal to the aqueous interface, necessitating orientation of the DLG molecule along the normal to the bilayer. Both carbonyl groups are H-bonded to H2O, but the secondary ester (sn-2) carbonyl is relatively more hydrated than the primary ester (sn-1) carbonyl. The 13C-carbonyl chemical shift data further suggest that the interfacial conformation resembles that of crystalline and liquid crystalline lamellar 1,2-dilauroyl-sn-glycero-3-phosphatidylethanolamine and certain PCs, in which the glycerol backbone is perpendicular to the bilayer plane. This conformation is different from that of crystalline 1,2-dilauroyl-sn-glycerol, in which the glycerol backbone is parallel to the bilayer plane. Between 1.5 and 8% DLG in vesicles, the chemical shift of the 1,2-DLG carbonyl at a given temperature was constant. However, above 8% DLG the chemical shift at each temperature increased with increasing DLG concentration, suggesting increased hydration at higher DLG content. At low temperatures 13C NMR spectra of vesicles with the highest proportions of 1,2-DLG studied (15 and 20%) showed two DLG carbonyl resonances, which most likely represent 1,2-DLG on outer and inner leaflets of the vesicle bilayer. The two peaks collapsed into a single resonance by 38 degrees C, at which temperature the two environments equilibrate with a rate constant of approximately 60 s-1 (t1/2 approximately 10 ms). Thus, transbilayer movement of DLG is extremely fast compared with phospholipids. In vesicles the 1,3-isomer of DLG exhibited a narrow carbonyl peak slightly downfield from that of 1,2-DLG. Acyl chain migration from 1,2-DLG to 1,3-DLG was monitored directly in the vesicle by time-dependent NMR measurements.  相似文献   

10.
An enzyme capable of dehalogenating vicinal haloalcohols to their corresponding epoxides was purified from the 3-chloro-1,2-propanediol-utilizing bacterium Arthrobacter sp. strain AD2. The inducible haloalcohol dehalogenase converted 1,3-dichloro-2-propanol, 3-chloro-1,2-propanediol, 1-chloro-2-propanol, and their brominated analogs, 2-bromoethanol, as well as chloroacetone and 1,3-dichloroacetone. The enzyme possessed no activity for epichlorohydrin (3-chloro-1,2-epoxypropane) or 2,3-dichloro-1-propanol. The dehalogenase had a broad pH optimum at about 8.5 and a temperature optimum of 50 degrees C. The enzyme followed Michaelis-Menten kinetics, and the Km values for 1,3-dichloro-2-propanol and 3-chloro-1,2-propanediol were 8.5 and 48 mM, respectively. Chloroacetic acid was a competitive inhibitor, with a Ki of 0.50 mM. A subunit molecular mass of 29 kDa was determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. With gel filtration, a molecular mass of 69 kDa was found, indicating that the native protein is a dimer. The amino acid composition and N-terminal amino acid sequence are given.  相似文献   

11.
A L Larrabee 《Biochemistry》1979,18(15):3321-3326
The results of transmission electron microscopic and ultracentrifugal studies of the size distributions of sonicated distearoylphosphatidylcholine vesicles are reported. Small vesicles (d approximately 300 A) were prepared by sonication of pure 1,2-distearoyl-3-sn-phosphatidylcholine in water and incubated at 4, 21, 40, 53, and 65 degrees C. The vesicle size distributions changed as a function of time at all temperatures below the phase-transition temperature but remained constant at the transition temperature and above. The sizes of structures to which the small vesicles are converted are the same at all temperatures, although the rates of conversion differ. The primary structures formed are identified as larger vesicles. The rate of loss of small vesicles is found to increase with decreasing temperature. At 4 and 21 degrees C small vesicles are converted to amorphous material, possibly irregular fragments of neat phase, in addition to being converted to larger vesicles. Trace amounts of an impurity commonly produced in the synthesis of 1,2-distearoyl-3-sn-phosphatidylcholine, 1,3-distearoyl-2-sn-phosphatidylcholine, are found to dramatically reduce the rate of loss of small vesicles at 21 degrees C.  相似文献   

12.
A high-yield synthesis of saturated, unsaturated, and short chain phosphatidylcholines from sn-glycero-3-phosphocholine is described. The procedure offers advantages over other reported procedures for the synthesis of phosphatidylcholine in that the large-scale synthesis and purification can be achieved in a minimum time. The procedure utilizes 4-pyrrolidinopyridine as a catalyst and moderate amounts of fatty acid anhydride (2 mol eq. of fatty acid anhydride per mol of OH) in a 1:1 mixture of benzene-dimethylsulfoxide (DMSO) at 40 degrees--42 degrees C (oilbath) for 2--5 hr. At the end of the reaction, the phosphatidylcholine can be purified in the usual manner or by using a Waters Prep LC/500 with a radially compressed silica gel column eluted with chloroform-methanol-water 60:30:4. At a flow rate of 200 ml/min, the phospholipid elutes in 10--15 min, depending on the chain length and unsaturation.  相似文献   

13.
The rearrangements for 2-phospha-4-silabicyclo[1.1.0]butane, analogous to the valence isomerization of the hydrocarbons bicyclobutane, 1,3-butadiene, and cyclobutene, were studied at the (U)QCISD(T)/6-311+G**//(U)QCISD/6-31G* level of theory. The monocyclic 1,2-dihydro-1,2-phosphasiletes are shown to be the thermodynamically preferred product, in contrast to the isomerization of the hydrocarbons, which favors the 1,3-butadiene structure. Furthermore, an unprecedented direct isomerization pathway to the 1,2-dihydro-1,2-phosphasiletes was identified. This pathway is competitive with the isomerization via the open-chain butadienes and becomes favorable when electron-donating substituents are present on silicon. Figure 2-Phospha-4-silabicyclo[1.1.0]butane can isomerize directly into the more stable P,Si-cyclobutene via an unprecedented [sigma2s+sigma2a] process, which becomes favorable over the isomerization via the P,Si-butadiene when electron-donating substituents are present on silicon.  相似文献   

14.
Structural analysis of the phosphomannan isolated from yeast-form cells of a pathogenic yeast, Candida albicans J-1012 strain, was conducted. Treatment of this phosphomannan (Fr. J) with 10 mM HCl at 100 degrees C for 60 min gave a mixture of beta-1,2-linked manno-oligosaccharides, from tetraose to biose plus mannose, and an acid-stable mannan moiety (Fr. J-a), which was then acetolyzed by means of an acetolysis medium, 100:100:1 (v/v) mixture of (CH3CO)2O, CH3COOH, and H2SO4, at 40 degrees C for 36 h in order to avoid cleavage of the beta-1,2 linkage. The resultant manno-oligosaccharide mixture was fractionated on a column of Bio-Gel P-2 to yield insufficiently resolved manno-oligosaccharide fractions higher than pentaose and lower manno-oligosaccharides ranging from tetraose to biose plus mannose. The higher manno-oligosaccharide fraction was then digested with the Arthrobacter GJM-1 alpha-mannosidase in order to cleave the enzyme-susceptible alpha-1,2 and alpha-1,3 linkages, leaving manno-oligosaccharides containing the beta-1,2 linkage at their nonreducing terminal sites, Manp beta 1----2Manp alpha 1----2Manp alpha 1----2Manp alpha 1----2Man, Manp beta 1----2Manp beta 1----2Manp alpha 1----2Manp alpha 1---- 2Manp alpha 1----2Man, and Manp beta 1----2Manp beta 1----2Manp beta 1----2Manp alpha 1---- 2Manp alpha 1----2Manp alpha 1----2Man. However, the result of acetolysis of Fr. J-a by means of a 10:10:1 (v/v) mixture of (CH3CO)2O, CH3COOH, and H2SO4 at 40 degrees C for 13 h was significantly different from that obtained by the mild acetolysis method; i.e., the amount of mannose was apparently larger than that formed by the mild acetolysis method. In summary, a chemical structure for Fr. J as a highly branched mannan containing 14 different branching moieties was proposed.  相似文献   

15.
Solvent mixtures containing trimethyl borate virtually eliminated the pronounced interconversion of 1,2- and 1,3-dipalmitins during their resolution by thin-layer chromatography on Silica Gel G. With trimethyl borate, an average of 1-2% of 1,2-dipalmitin was converted to 1,3-dipalmitin. A four-directional-development TLC procedure incorporating trimethyl borate resolves cholesteryl glucoside, ceramides, monogalactosyl diglyceride, 1- and 2-monopalmitin, palmitic acid, cholesterol, 1,2- and 1,3-dipalmitin, tripalmitin, methyl palmitate, cholesteryl palmitate, beta-carotene and some of its degradation products, squalene, and tetracosane. Digalactosyl diglyceride, phosphatidic acid, phosphatidylglucose, cerebrosides, and other phospholipids remain near the origin. A mixture containing triolein, 1,2- and 1,3-diolein, 1- and 2- monoolein, oleic acid, and cholesterol was resolved in one dimension. A similar series of palmitic-containing neutral lipids was also resolvable in one dimension. These procedures were applied to the TLC of human sera lipids.  相似文献   

16.
Incubation of glyceryl trinitrate (GTN) with 5% (w/v) rat brain homogenate (RBH) resulted in biotransformation of the organic nitrate vasodilator drug to a mixture of glyceryl-1,2-dinitrate (1,2-GDN) and glyceryl-1,3-dinitrate (1,3-GDN). Heating of the RBH at 100 degrees C for 5 min and (or) pretreatment with 5 mM N-ethylmaleimide at 37 degrees C for 10 min demonstrated that about two-thirds of the GTN biotransformation activity was due to a sulfhydryl-dependent enzymatic process resulting in the predominant formation of 1,2-GDN, and that the remaining biotransformation activity was due to a sulfhydryl-dependent nonenzymatic process resulting in the selective formation of 1,3-GDN. In a preliminary experiment, nitric oxide formation was observed during the incubation of GTN with RBH under anaerobic conditions. These data support the idea that some of the therapeutic and adverse effects of GTN are mediated through its action in the central nervous system.  相似文献   

17.
We have developed a liquid chromatographic-mass spectrometric method for the simultaneous determination of nitroglycerin (NTG) and its active metabolites, glyceryl 1,2-dinitrate (1,2-GDN) and glyceryl 1,3-dinitrate (1,3-GDN), for metabolism studies in cell cultures. 1,2,4-Butanetriol-1,4-dinitrate was chosen as an internal standard. Using a linear gradient of water/methanol containing 0.025 mM NH(4)Cl, the compounds were eluted within 12.5 min on an Allure Aqueous C(18) column (100 mm x 2.1 mm). Detection and quantification was achieved with multiple reaction monitoring in the negative ion mode. Intra- and inter-day variabilities for simultaneous determination of the three nitrates were below 10 and 18%, respectively, over a range of NTG and GDN concentrations of 0.5-15 ng/ml. The lower limit of quantification was found to be about 0.01 ng on column. Application of this method was illustrated through in vitro metabolism studies of NTG in culture media bathing LLC-PK1 cells and human vascular smooth muscle cells (HA-VSMC) at 37 degrees C. The degradation half-life of NTG was found to be 4.5 +/- 0.4 h and 39.2 +/- 0.02 h, respectively, for LLC-PK1 cells versus HA-VSMC. At 5 h, the 1,2-GDN versus 1,3-GDN metabolite distribution ratio in the bathing medium was found to be 1.5 +/- 0.1 and 0.2 +/- 0.02 for LLC-PK1 and HA-VSMC cells, respectively. With this method, the degradation half-life of NTG in rat plasma at 37 degrees C was shown to be 26.8 +/- 1.8 min, consistent with previous values obtained using gas chromatography.  相似文献   

18.
Acyl-coenzyme A:cholesterol acyltransferase (ACAT) assays are usually performed by incubation of the enzyme with a labeled substrate followed by thin-layer chromatography separation and subsequent quantification of cholesteryl esters (CE) formed. Herein, a method is described for rapid separation of CE from other lipids, by elution from a silica gel column with a solvent mixture of petroleum ether/diethyl ether (98:2, v/v). Silica gel column chromatography is reliable and more rapid and safer than TLC. The best results were obtained when the reaction was stopped by Dole extraction followed by CE separation on a silica gel column. Assays for ACAT from rat intestinal microsomes showed that the specific activity values obtained using this method were reproducible and in good agreement with those obtained by conventional TLC method.  相似文献   

19.
A Lange  D Marsh  K H Wassmer  P Meier  G Kothe 《Biochemistry》1985,24(16):4383-4392
The electron spin resonance spectra of the 1-myristoyl-2-[6-(4,4-dimethyloxazolidine-N-oxyl)myristoyl]-sn-glycero- 3-phosphocholine spin-label in highly oriented, fully hydrated bilayers of 1,2-dimyristoyl-sn-glycero-3-phosphocholine have been studied as a function of temperature and magnetic field orientation. The oriented spectra show clear indications of slow motional components (rotational correlation times greater than 3 ns) even in the fluid phase (T greater than 23 degrees C), indicating that motional narrowing theory is not applicable to the spectral analysis. The spectra have been simulated by a comprehensive line-shape model that incorporates trans-gauche isomerization in addition to restricted anisotropic motion of the lipid long molecular axis and that is valid in all motional regimes. In the gel (L beta') phase the spin-label chains are found to be tilted at 28 degrees with respect to the normal of the orienting plane. In the intermediate (P beta') phase there is a continuous distribution of tilt angles between 0 degrees and 25 degrees. In fluid (L alpha) phase there is no net tilt of the lipid chains. The chains rotate at an intermediate rate about their long axis in the fluid phase (tau R,parallel = 1.4-6.6 ns for T = 50-25 degrees C), but the reorientation of the chain axis is much slower (tau R, perpendicular= 13-61 ns for T = 50-25 degrees C), whereas trans-gauche isomerization (at the C-6 position) is rapid (tau J less than or equal to 0.2 ns). Below the chain melting transition both chain reorientation and chain rotation are at the ESR rigid limit (tau R greater than or equal to 100 ns), and trans-gauche isomerization is in the slow-motion regime (tau J = 3.7-9.5 ns for T = 22-2 degrees C). The chain order parameter increases continuously with decreasing temperature in the fluid phase (SZZ = 0.47-0.61 for T = 50-25 degrees C), increases abruptly on going below the chain melting transition, and then increases continuously in the intermediate phase (SZZ = 0.79-0.85 for T = 22-14 degrees C) to an approximately constant value in the gel phase (SZZ congruent to 0.86 for T = 10-2 degrees C).(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

20.
The physical properties of hydrated multilamellar sample of 1,2-dimyristamido-1,2-deoxyphosphatidylcholine (DDPC) were investigated by means of differential scanning calorimetry (DSC), static X-ray diffraction, and simultaneous DSC and X-ray diffraction. The DDPC is a synthetic sphingomyelin analogue and has two amide bonds in its hydrophobic parts. This paper reports on metastable phase behavior of the hydrated DDPC sample. By cooling from a chain-melted state at the rates of greater than 4 degrees C min(-1), hydrated DDPC bilayers form a metastable gel phase. In the gel phase, the hydrophobic chains are tilted with respect to the bilayer normal, as like the gel phase of glycero-phosphatidylcholines. By heating, the metastable gel phase is transformed in to a stable phase associated with an exothermic heat event at 18.3 degrees C (DeltaH=14.6 kJ mol(-1)) and then the stable phase is transformed into a liquid-crystalline phase at 25.6 degrees C (DeltaH=42 kJ mol(-1)). The incubation at 17 degrees C for more than 1 h also induces the formation of the stable phase. In the stable phase, the hydrophobic chains are packed into highly ordered crystal-like structure. However, the X-ray diffraction pattern of the stable phase suggested that the entire DDPC molecules do not form a two-dimensional molecular ordered lattice, differing from normal subgel phase of glycero-phosphatidylcholines. The structure and phase behavior of DDPC revealed by the present study are discussed from the viewpoint of hydrogen bonds.  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