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.
B Tümmler  U Herrmann  G Maass  H Eibl 《Biochemistry》1984,23(18):4068-4074
The thermodynamics and kinetics of the subtransition L epsilon----P beta of sonicated unilamellar vesicles of 1-myristoyl-2-stearoylphosphatidylcholine (1M-2S-PC) and of 1-stearoyl-2-myristoylphosphatidylcholine (1S-2M-PC) were studied by equilibrium cooling curves and temperature-jump relaxation spectrometry with an anthracenophane cryptand as a mobile fluorescent probe. The unilamellar vesicles exhibit the midpoint temperature TsII of the subtransition about 10 degrees C below the respective main transition. The kinetics of the subtransition in the time range between 10(-4) and 10(3) s is characterized by a cooperative relaxation process in the range of milliseconds and a further noncooperative process in the range of seconds. The slow process is assigned to the rearrangement of lattice defects. The fast process is evaluated in terms of a cyclic reaction scheme that consists of two pathways for the biomolecular association of probe and vesicle coupled with the conformational change of the lipid matrix during the subtransition. The analysis reveals that the fast process comprises the nucleation and growth of cluster. The cooperative lattice transformation of the subtransition follows a first-order rate law. The rate constants at TsII are 70 s-1 for 1S-2M-PC and 170 s-1 for 1M-2S-PC. Since the plots of the relaxation time vs. the degree of transition are in accordance with the predictions of the linear Ising model, it is concluded that clusters are propagated anisotropically in a linear fashion; e.g., fluidlike P beta conformations grow along the ripple.  相似文献   

2.
《FEBS letters》1987,224(2):283-286
The ‘main’ phase transition Lβ→Lα of hydrated 1,2-dipalmitoylphosphatidylethanolamine (DPPE) bilayers in excess water affects the ESR order parameter S33 of N-cetyl-N,N-dimethyl-N-tempoylammonium bromide (CAT-16), 5-doxylstearic acid (5-DSA) and 16-doxylstearic acid (16-DSA) spin probes. The ‘pretransition’ and ‘subtransition’ suggested to occur in hydrated DPPE by Chowdhry et al. [(1984) Biophys. J. 45, 901–904] and Silvius et al. [(1986) Biochemistry 25, 4249–4258], respectively, affect exclusively the S33 of CAT-16, but not that of 5-DSA and 16-DSA spin probes. The subtransition occurs about 15 ± 1°C below the main transition.  相似文献   

3.
dl-Dipalmitoylphosphatidylcholine multilamellar vesicle suspensions were examined by the method of differential scanning calorimetry. A lack of the subtransition at 18°C was established. Such a subtransition is characteristic for l-dipalmitoylphosphatidylcholine suspensions. This lack is supposed to be the result of the impossibility of the racemic phospholipid mixture to form the low-temperature crystal structure Lc.  相似文献   

4.
Suspensions of dipalmitoylphosphatidylglycerol (DPPG) have been analyzed by differential scanning calorimetry, equilibrium and differential scanning dilatometry, and X-ray diffraction techniques. After the DPPG suspensions are stored several days at 2 degrees C, a new phase transition is observed at a lower temperature than either the main transition or the pretransition. This subtransition has an enthalpy of about 6 kcal/mol and occurs at about 20 degrees C, the exact temperature depending on the buffer used. The lipid partial specific volume increases by 0.035 mL/g upon warming through the subtransition. X-ray diffraction patterns from suspensions in the subgel phase contain orders of a lamellar repeat and several additional sharp and broad wide-angle reflections between 8 and 2 A. As the water content in the specimen is reduced, the lamellar repeat period decreases, whereas the spacings and intensities of these additional wide-angle reflections are unchanged. These data indicate that on incubation at 2 degrees C the lipid molecules crystallize in the plane of each bilayer. X-ray experiments also show that this subgel phase converts to the normal L beta' gel phase above the subtransition.  相似文献   

5.
The kinetics of the hydrogen-deuterium exchange reactions of double-helical poly (rI) · poly (rC), single-stranded poly(rC) and poly(rI), inosine, and cytosine- 5′-phosphoric acid have been examined, at various temperatures in the range 20 °C to 52 °C, by stopped-flow ultraviolet spectrophotometry, in the region 270 to 300 nm. For the solution of double-helical poly(rI) · poly(rC), two first-order deuteration reactions were found: a fast one and a slow one. At 25 °C and at pH 7.0, the rate constant was 12.3 s?1 for the fast reaction, and 0.13 s?1 for the slow reaction. The rate constant of the fast reaction is nearly equal to that of the single-stranded poly(rC) (12.6 s?1), and is assigned to the deuteration at the amino hydrogen (that is, free from the C · I hydrogen bond) of the cytosine residue. The slow reaction is attributable to the deuteration of the two hydrogens: the amino hydrogen of rC and imide hydrogen of rI, which are rapidly exchanging with each other within every rC · rI base-pair. From the observed temperature effect on this slow reaction rate, it has been concluded that there are two types of “opening process” that are relevant to the hydrogen exchange reaction; one of them is predominent in the range 47 °C to 52 °C and the other in the temperature region lower than 47 °C. The enthalpy (H) and entropy (S) differences of the “open” and “closed” forms in the former type process are ΔH = 167 kcal per mole and ΔS = 507 e.u., while in the latter ΔH = 8.1 kcal per mole and ΔS = 10 e.u..  相似文献   

6.
The syntheses of the tetraazamacrocyclic ligands 1,4,7,11-tetraazacyclotetradecane (isocyclam) and 1,5,9,13-tetraazacyclohexadecane ([16]aneN4) in two steps starting from the corresponding tetraamine and diethylmalonate is reported. The trans-dicyanochromium(III) complexes, trans-[Cr(isocyclam)(CN)2]PF6 and trans-[Cr([16]aneN4)(CN)2]PF6 have also been prepared. Both are 2Eg emitters with 0-0 band emission wavelengths at 721.2 and 704.8 nm, respectively. The isocyclam complex has a room temperature excited state lifetime of 147 μs in aqueous solution which increases to 215 μs upon macrocyclic N-H deuteration, whereas the corresponding lifetime of the [16]aneN4 complex is 25 μs and is unaffected by macrocyclic N-H deuteration. The implications of the temperature dependence of the excited state lifetimes are also presented.  相似文献   

7.
Effect of propylgallate (PrG) on the thermotropic behavior of mixtures of dipalmitoylphosphatidylglycerol (DPPG) and Ca2+ was studied by means of differential scanning calorimetry (DSC). In the case of DPPG or DPPG/Ca (molar ratio, 15 : 1), the transition temperature (Tm) of the main transition and the subtransition decreased from 40 degrees C to 29 degrees C and from 29 degrees C to 20 degrees C, respectively, with an increase in the concentration of PrG. The addition of PrG to the DPPG/Ca mixture induced a shoulder on the high temperature side in the reheating scan. Neither PrG nor low concentrations of Ca2+ bind to the Lc phase of DPPG. When the molar ratio of DPPG to Ca was 1 : 1, the subtransition did not occur, that is, only the main transition (Tm = 90 degrees C) appeared. The Tm of the main transition was slightly affected by PrG. On the addition of PrG, another metastable endothermic transition peak (Tm = 78 degrees C) appeared. It is concluded that Ca2+ and PrG inhibit each other's binding.  相似文献   

8.
Dilatometric studies of the subtransition in dipalmitoylphosphatidylcholine   总被引:4,自引:0,他引:4  
J F Nagle  D A Wilkinson 《Biochemistry》1982,21(16):3817-3821
The phase transition between the newly discovered low-temperature subgel phase and the gel phase of dipalmitoylphosphatidylcholine has been studied by using dilatometry. Equilibrium measurements show that the subtransition upon heating is centered at 13.5 degrees C, has a dilatometric half-width of 0.6 degree C, and comprises a specific volume change of 0.009 mL/g (about one-fourth the size of the main transition). When the gel phase is cooled, the subtransition does not occur until below 5 degrees C. The rate of formation as a function of incubation temperature for 1 degree C less than TI less than 6 degrees C was determined; it is not well fit by quantitative theories based upon homogeneous nucleation. However, some form of nucleation is present since temperature-jump studies show that once the subgel phase has started to form, it continues to grow in the range 6 degrees C less than TJ less than 12.8 degrees C. Thus, the true transition temperature lies between 12.8 and 13.5 degrees C, but nucleation of the subgel phase is severely retarded above 6 degrees C, leading to the large hysteresis observed upon cooling.  相似文献   

9.
The mechanism of the subtransitions (Lc to L beta') in L-dipalmitoylphosphatidylcholine bilayers in excess water has been investigated by time-resolved X-ray diffraction using synchrotron radiation. The temperature dependence of the diffraction patterns closely correlate with the asymmetric excess specific heat variation recorded by differential scanning calorimetry. During the subtransition two prominent wide-angle reflections, characteristic of the low-temperature crystalline phase, Lc, gradually change such that a sharp peak at a spacing of 0.430 nm decreases in intensity and ultimately disappears while a broader peak initially located at 0.375 nm progressively shifts to an eventual spacing of 0.410 nm. This behaviour is interpreted as a lateral deformation of the acyl chain packing subcell as the chains begin to rotate until a state is reached where the chains pack on a regular hexagonal array characteristic of the L beta phase. An increase in lamellar repeat distance from 6.0 to 6.4 nm takes place simultaneously with the acyl chain rearrangement at relatively low (5 K/min) as well as high (6 K/s) heating rates. As judged from the shape of the wide-angle peak, transformation to L beta' phase occurs some minutes after transition to the L beta phase. The X-ray data characterise the subtransition as a continuous (second order) phase transition in which a presumably orthorhombic subcell is transformed into a hexagonal subcell in a gradual process. In temperature jump experiments at 6 K/s between 0 degree C and 80 degrees C the relaxation time of the subtransition was found to be about 5 s while the relaxation time of the main gel to liquid-crystalline transition was about 2 s.  相似文献   

10.
A statistical mechanical model with experimentally proved facts as starting points is presented. This model explains on molecular level, the pre- and subtransitions appearing in lipid membranes. The model describes the main features of the transitions, the hysteresis of the subtransition and the mobility changes of the heads and chains at these transitions. The model was expanded for phosphatidylcholine homologues with arbitrary chain lengths, and a qualitative agreement in the case of pretransition as far as a quantitative one for the subtransition were found.  相似文献   

11.
Phosphatidylcholine (PC) is a major component of eukaryotic cell membranes and one of the most commonly used phospholipids for reconstitution of membrane proteins into carrier systems such as lipid vesicles, micelles and nanodiscs. Selectively deuterated versions of this lipid have many applications, especially in structural studies using techniques such as NMR, neutron reflectivity and small-angle neutron scattering. Here we present a comprehensive study of selective deuteration of phosphatidylcholine through biosynthesis in a genetically modified strain of Escherichia coli. By carefully tuning the deuteration level in E. coli growth media and varying the deuteration of supplemented carbon sources, we show that it is possible to achieve a controlled deuteration for three distinct parts of the PC lipid molecule, namely the (a) lipid head group, (b) glycerol backbone and (c) fatty acyl tail. This biosynthetic approach paves the way for the synthesis of specifically deuterated, physiologically relevant phospholipid species which remain difficult to obtain through standard chemical synthesis.  相似文献   

12.
F Schaeffer  A Kolb    H Buc 《The EMBO journal》1982,1(1):99-105
To understand the denaturation process of short DNA segments we have chosen a 203-base pair (bp) restriction fragment containing the lactose control region. A steady decrease in GC content exists between its i proximal and z proximal ends. We confirm that this fragment melts at low salt in two subtransitions. A GC to AT mutation in the AT-rich region (mutation UV5) increases the number of denatured base pairs in the first subtransition and decreases the cooperativity of the melting process. A GC to AT mutation in the GC-rich region (mutation L8) decreases the number of denatured base pairs in the first subtransition and increases the cooperativity. These mutations induce the same shift in the temperature of half denaturation. The effects of both mutations are additive. A short deletion at the z end of the fragment affects only the first subtransition. When four GC pairs are added to both end, the fragment melts in one transition. Comparison with the results obtained with a larger 789-bp lac fragment reveals strong end effects on base pair stability and suggests that denaturation of the 203-bp fragment proceeds unidirectionally from the z end. Good agreement is shown with the predictions made with the "z ipper model" of Crothers et al. (1965).  相似文献   

13.
The tension-driven gating transition in the large mechanosensitive channel MscL proceeds through detectable states of intermediate conductance. Gain-of-function (GOF) mutants with polar or charged substitutions in the main hydrophobic gate display altered patterns of subconducting states, providing valuable information about gating intermediates. Here we present thermodynamic analysis of several GOF mutants to clarify the nature and position of low-conducting conformations in the transition pathway. Unlike wild-type (WT) MscL, which predominantly occupies the closed and fully open states with very brief substates, the mild V23T GOF mutant frequently visits a multitude of short-lived subconducting states. Severe mutants V23D and G22N open in sequence: closed (C) --> low-conducting substate (S) --> open (O), with the first subtransition occurring at lower tensions. Analyses of equilibrium state occupancies as functions of membrane tension show that the C-->S subtransition in WT MscL is associated with only a minor conductance increment, but the largest in-plane expansion and free energy change. The GOF substitutions strongly affect the first subtransition by reducing area ((Delta)A) and energy ((Delta)E) changes between C and S states commensurably with the severity of mutation. GOF mutants also exhibited a considerably larger (Delta)E associated with the second (S-->O) subtransition, but a (Delta)A similar to WT. The area changes indicate that closed conformations of GOF mutants are physically preexpanded. The tension dependencies of rate constants for channel closure (k(off)) predict different positions of rate-limiting barriers on the energy-area profiles for WT and GOF MscL. The data support the two-gate mechanism in which the first subtransition (C-->S) can be viewed as opening of the central (M1) gate, resulting in an expanded water-filled "leaky" conformation. Strong facilitation of this step by polar GOF substitutions suggests that separation of M1 helices associated with hydration of the pore in WT MscL is the major energetic barrier for opening. Mutants with a stabilized S1 gate demonstrate impeded transitions from low-conducting substates to the fully open state, whereas extensions of S1-M1 linkers result in a much higher probability of reverse O-->S transitions. These data strongly suggest that the bulk of conductance gain in the second subtransition (S-->O) occurs through the opening of the NH2-terminal (S1) gate and the linkers are coupling elements between the M1 and S1 gates.  相似文献   

14.
The two dihydroflavonols, 3,5,4′-trihydroxy-7-methoxyflavanone and 3,5,4′-trihydroxy-7,3′-dimethoxyflavanone, and naringenin were isolated from aerial parts of Artemisia dracunculus. The mass spectrum of 3,5,4′-trideuteroxy-7-methoxyflavanone is described as an example of the usefulness of deuteration in the analysis of certain flavonoids.  相似文献   

15.
The H/D exchange catalysis using the Ir(I) complex [TpMe2Ir(η4-2,3-dimethylbutadiene)] (TpMe2=hydridotris(3,5-dimethylpyrazolyl)borate) as the precatalyst was studied for selective deuteration of norbornene derivatives. In dependence of the norbornene substitution in 2,3 positions, selective deuteration of the norbornene double bond could be achieved. (±)-endo,exo-6-Deutero-bicyclo[2.2.1]hept-5-ene-2,3-dicarboxylic acid diethyl ester was isolated in 82% yield.  相似文献   

16.
Mixtures of 1,2-dipalmitoyl- and 1,2-O-dihexadecyl-sn-glycero-3-phosphocholine (DPPC and DHPC) in dispersion with excess water were studied by differential scanning calorimetry (DSC) and X-ray diffraction techniques. The transition parameters of the main gel-to-liquid crystalline transition show a monotonous dependence on the composition, indicating ideal miscibility of the two lipids, in keeping with the closely similar structures of the pure, hydrated lipids in the P beta' and L alpha states. The pre-transition shows a depression to a minimum temperature of 23 degrees C occurring around equimolar mixtures. Below the pre-transition temperatures, the L beta' gel phase of DPPC maintains bimolecular structure up to DHPC admixtures of 50 mol%, with adaptations in hydrocarbon chain packing and multilayer periodicity. On the side of DHPC, the interdigitated gel structure shows full solubility for DPPC up to equimolarity without major structural changes. The crystalline Lc-phase of DPPC exhibits immiscibility with DHPC, demonstrated by the fact that the subtransition is abolished already at less than 15 mol% DHPC. DHPC, below its subtransition, can accommodate up to 50 mol% DPPC within an interdigitated layer structure with unperturbed, crystalline hydrocarbon chain packing.  相似文献   

17.
The effects of deuteration and of changes in solvent composition on the thermo dynamics of the helix–coil transition have been studied by calorimetric and optical measurements in the poly-γ-benzyl-L -glutamate–dichloroacetic acid–1,2-dichloro-ethanc system. For a given solvent composition, deuteration of the polypeptide and of the acid lowers the transition temperature Tc, while an increase in the volume fraction of acid in the solvent raises Tc. A rise in Tc is accompanied by a decrease in both the van't Hoff and the calorimetric heats of transition, but at different rates. The result is a temperature dependency in the Zimm-Bragg cooperation parameter σ. Possible causes of this result and its implications are discussed.  相似文献   

18.
We have investigated the effect of deuteration of non‐exchangeable protons on protein global thermal stability, hydrophobicity, and local flexibility using well‐known thermostable model systems such as the villin headpiece subdomain (HP36) and the third immunoglobulin G‐binding domain of protein G (GB3). Reversed‐phase high‐performance liquid chromatography (RP‐HPLC) measurements as a function of temperature probe global thermal stability in the presence of acetonitrile, while differential scanning calorimetry determines thermal stability in solution. Both indicate small but measurable changes in the order of several degrees. RP‐HPLC also permitted quantification of the effect of deuteration of just three core phenylalanine side chains of HP36. NMR dynamics investigation has focused on methyl axes motions using cross‐correlated relaxation measurements. The analysis of order parameters provided a complex picture indicating that deuteration generally increases motional amplitudes of sub‐nanosecond motion in GB3 but decreases those in HP36. Combined with earlier dynamics measurements at Cα–Cβ sites and backbone sites of GB3, which probed slower time scales, the results point to the need to probe multiple atoms in the protein and variety of time scales to the discern the full complexity of the effects of deuteration on dynamics.  相似文献   

19.
Calorimetric, X-ray diffraction, and 31P nuclear magnetic resonance (NMR) studies of aqueous dispersions of 1,2-dihexadecyl-sn-glycero-3-phosphocholine (DHPC) gel phases at low temperatures (-60 to 22 degrees C) show thermal, structural, and dynamic differences when compared to aqueous dispersions of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) gel phases at corresponding temperatures. Differential scanning calorimetry of DHPC dispersions demonstrates a reversible, low-enthalpy "subtransition" at 4 degrees C in contrast to the conditionally reversible, high-enthalpy subtransition observed at 17 degrees C for annealed DPPC bilayers. X-ray diffraction studies indicate that DHPC dispersions form a lamellar gel phase with dav congruent to 46 A both above and below the "subtransition". It is suggested that the reduced dav observed for DHPC (46 A as compared to 64 A in DPPC) is due to an interdigitated lamellar gel phase which exists at all temperatures below the pretransition at 35 degrees C. 31P NMR spectra of DHPC gel-phase bilayers show an axially symmetric chemical shift anisotropy powder pattern which remains sharp down to -20 degrees C, suggesting the presence of fast axial diffusion. In contrast, 31P spectra of DPPC bilayers indicate this type of motion is frozen out at approximately 0 degrees C.  相似文献   

20.
The mobility of 5-doxyl stearic acid spin label (5-SASL) in the gel phase of dipalmitoylphosphatidylcholine membranes between the main transition and subtransition temperatures was studied as a function of cholesterol content. Very small amounts of cholesterol (0.01-1 mol%) cause a dramatic increase in the mobility of 5-SASL. Temperature-drop experiments from 38 degrees C to 28 degrees C were made across the pretransition temperature and the rate of approach to equilibrium was measured. Cholesterol at low concentrations also affects this rate. The membrane reached equilibrium after 10 h in the absence of cholesterol, 3 h at 0.01 mol% cholesterol, and less than 10 min at 0.03 mol% cholesterol.  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