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.
Redox components bound with chromatophores of Rhodospirillum rubrum, and pure samples of ubiquinone-10 and rhodoquinone were studied polarographically at 24 degrees. In a mixture of ethanol and water (4 : 1, v/v) at pH 7, ubiquinone-10 and rhodoquinone had half-wave potentials (E1/2) OF +43 MV and -63 mV, respectively. For both quinones, values of the electron transfer number (n) were 2 , and plots of E1/2 versus pH formed straight lines with slopes of -30 mV/pH in the neutral pH range; thus, values of the proton transfer number (n-a) were estimated to be 1 for both quinones. When bound with chromatophores, ubiquinone-10 and rhodoquinone had E1/2 values of +50 mV (n=2) and -30 mV (n=2), respectively, at pH 7. Values of (n-a) were estimated to be 1 for ubiquinone-10 and 2 for rhodoquinone. A component (POC-170) thought to be one of the active center bacteriochlorophylls (Liac-890) was characterized; it has E1/2 value of -170 mV at pH 7 and its oxidation-reduction is possibly brought about by dehydrogenation-hydrogenation. Conceivably, the oxidation-reduction sites of ubiquinone-10, rhodoquinone and POC-170 partly, if not all, exist on the surface of chromatophore membrane or project outside the membrane, because of their accessibility to the polarographic electrode.  相似文献   

2.
The location of ubiquinone-10 in phospholipid bilayers was analyzed using a variety of physical techniques. Specifically, we examined the hypothesis that ubiquinone localizes at the geometric center of phospholipid bilayers. Light microscopy of dipalmitoylphosphatidylcholine at room temperature in the presence of 0.05-0.5 mol fraction ubiquinone showed two separate phases, one birefringent lamellar phase and one phase that consisted of isotropic liquid droplets. The isotropic phase had a distinct yellow color, characteristic of melted ubiquinone. [13C]NMR spectroscopy of phosphatidylcholine liposomes containing added ubiquinone indicated a marked effect on the 13C-spin lattice relaxation times of the lipid hydrocarbon chain atoms near the polar head region of the bilayer, but almost no effect on those atoms nearest the center of the bilayer. X-ray diffraction experiments showed that for phosphatidylcholine bilayers, both in the gel and liquid-crystal-line phases, the presence of ubiquinone did not change either the lamellar repeat period or the wide-angle reflections from the lipid hydrocarbon chains. In electron micrographs, the hydrophobic freeze-fracture surfaces of bilayers in the rippled (P beta') phase were also unmodified by the presence of ubiquinone. These results indicate that the ubiquinone which does partition into the bilayer is not localized preferentially between the monolayers, and that an appreciable fraction of the ubiquinone forms a separate phase located outside the lipid bilayer.  相似文献   

3.
Bilayer structure and interbilayer repulsive pressure were measured from 5 to 50 degrees C by the osmotic stress/x-ray diffraction method for both gel and liquid crystalline phase lipid bilayers. For gel phase dibehenoylphosphatidylcholine (DBPC) the bilayer thickness and pressure-distance relations were nearly temperature-independent, and at full hydration the equilibrium fluid spacing increased approximately 1 A, from 10 A at 5 degrees C to 11 A at 50 degrees C. In contrast, for liquid crystalline phase egg phosphatidylcholine (EPC), the bilayer thickness, equilibrium fluid spacing, and pressure-distance relation were all markedly temperature-dependent. As the temperature was increased from 5 to 50 degrees C the EPC bilayer thickness decreased approximately 4 A, and the equilibrium fluid spacing increased from 14 to 21 A. Over this temperature range there was little change in the pressure-distance relation for fluid spacings less than approximately 10 A, but a substantial increase in the total pressure for fluid spacings greater than 10 A. These data show that for both gel and liquid crystalline bilayers there is a short-range repulsive pressure that is nearly temperature-independent, whereas for liquid crystalline bilayers there is also a longer-range pressure that increases with temperature. From analysis of the energetics of dehydration we argue that the temperature-independent short-range pressure is consistent with a hydration pressure due to polarization or electrostriction of water molecules by the phosphorylcholine moiety. For the liquid crystalline phase, the 7 A increase in equilibrium fluid spacing with increasing temperature can be predicted by an increase in the undulation pressure as a consequence of a temperature-dependent decrease in bilayer bending modulus.  相似文献   

4.
S Ali  D Zakim 《Biophysical journal》1993,65(1):101-105
The thermotropic properties of multilamellar vesicles of dimyristoylphosphatidylcholine (DMPC), dipalmitoylphosphatidylcholine (DPPC), and distearoylphosphatidylcholine (DSPC), as a function of the concentration of bilirubin in the range of 0.1 to 1 mol%, were measured. The exact effects of bilirubin depended on the chain length of the polymethylene chains. But the general effects of bilirubin were the same in all systems. At the lowest concentrations tested (0.1 mol bilirubin/100 mol phospholipid (0.1 mol%)), bilirubin broadened and shifted to higher temperatures the main phase transitions of all bilayers. For DPPC and DSPC, but not DMPC, this concentration of bilirubin was associated with a new transition at 25 degrees C (DPPC) or 34 degrees C (DSPC). Bilirubin at 0.2 mol% was required for the detection of a similar transition (at 13.7 degrees C) in DMPC. Higher concentrations of bilirubin (> 0.2 mol%) suppressed completely the main phase transitions in all bilayers but increased the enthalpy of the new transition. Maximal values of delta H for these transitions were reached at 0.5, 0.25, and 0.2 mol% bilirubin in DMPC, DPPC, and DSPC, respectively. Values of delta H and delta S for these transitions were far larger than for the corresponding gel-to-liquid crystal transitions in pure lipid bilayers but were equal to those expected for a transition between crystalline and liquid crystalline phases.  相似文献   

5.
A G Lee 《Biochemistry》1977,16(5):835-841
The partitioning of the spin label 2,2,6,6-tetramethylpiperidinyl-1-oxy (Tempo) into phosphatidylcholine bilayers and the monomer-aggregate equilibrium for chlorophyll a incorporated into phosphatidylcholine bilayers have been interpreted in terms of the formation of defects in the gel-phase lipid, starting some 20 degrees C below the temperature of the main gel to liquid crystalline phase transition. By contrast, defects seem to be largely absent from bilayers of dipalmitoylphosphatidylethanolamine in the gel phase. The defect structure accounts for the continuous nature of the phase transition for phosphatidylcholines, and also for the increase in width of the transition caused by the addition of alcohols.  相似文献   

6.
X Peng  J Jonas 《Biochemistry》1992,31(28):6383-6390
High-pressure 31P NMR was used for the first time to investigate the effects of pressure on the structure and dynamics of the phosphocholine headgroup in pure 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) multilamellar aqueous dispersions and in DPPC bilayers containing the positively charged form of the local anesthetic tetracaine (TTC). The 31P chemical shift anisotropies, delta sigma, and the 31P spin-lattice relaxation times, T1, were measured as a function of pressure from 1 bar to 5 kbar at 50 degrees C for both pure DPPC and DPPC/TTC bilayers. This pressure range permitted us to explore the rich phase behavior of DPPC from the liquid-crystalline (LC) phase through various gel phases such as gel I (P beta'), gel II (L beta'), gel III, gel IV, gel X, and the interdigitated, Gi, gel phase. For pure DPPC bilayers, pressure had an ordering effect on the phospholipid headgroup within the same phase and induced an interdigitated Gi gel phase which was formed between the gel I (P beta') and gel II (L beta') phases. The 31P spin-lattice relaxation time measurements showed that the main phase transition (LC to gel I) was accompanied by the transition between the fast and slow correlation time regimes. Axially symmetric 31P NMR lineshapes were observed at pressures up to approximately 3 kbar but changed to characteristic axially asymmetric rigid lattice lineshapes at higher pressures (3.1-5.1 kbar).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
The location of ubiquinone-10 in phospholipid bilayers was analyzed using a variety of physical techniques. Specifically, we examined the hypothesis that ubiquinone localizes at the geometric center of phospholipid bilayers. Light microscopy of dipalmitoylphosphatidylcholine at room temperature in the presence of 0.05–0.5 mol fraction ubiquinone showed two separate phases, one birefringent lamellar phase and one phase that consisted of isotropic liquid droplets. The isotropic phase had a distinct yellow color, characteristic of melted ubiquinone. [13C]NMR spectroscopy of phosphatidylcholine liposomes containing added ubiquinone indicated a marked effect on the 13C-spin lattice relaxation times of the lipid hydrocarbon chain atoms near the polar head region of the bilayer, but almost no effect on those atoms nearest the center of the bilayer. X-ray diffraction experiments showed that for phosphatidylcholine bilayers, both in the gel and liquid-crystal-line phases, the presence of ubiquinone did not change either the lamellar repeat period or the wide-angle reflections from the lipid hydrocarbon chains. In electron micrographs, the hydrophobic freeze-fracture surfaces of bilayers in the rippled (Pβ′) phase were also unmodified by the presence of ubiquinone. These results indicate that the ubiquinone which does partition into the bilayer is not localized preferentially between the monolayers, and that an appreciable fraction of the ubiquinone forms a separate phase located outside the lipid bilayer.  相似文献   

8.
(1) The 129 MHZ and 36.4 MHZ 31 P NMR spectra of unsonicated liposomes consisting of phosphatidylcholines of varying chain length and unsaturation have been investigated. (2) In the liquid crystalline state the 31 P NMR liposome spectra are similar for both saturated and unsaturated phosphatidylcholines, demonstrating that the motion of the polar headgroup is not sensitive to the fatty acid composition in the disordered liquid crystalline state. (3) Below the hydrocarbon phase transition temperature there is a marked increase in the linewidth of the 31P NMR liposome spectra, indicating a reduction in the motion of the polar headgroup. (4) The addition of equimolar concentrations of cholesterol to phosphatidylcholine eliminates phase transition effects experienced by the polar headgroup. The motion of the polar headgroup is then very similar to that obtained in the liquid crystalline state for pure phosphatidylcholine bilayers. (5) In the liquid crystalline state the motion of the polar headgroup in the phosphate region is insensitive to changes in the available area per phosphatidy-choline molecule.  相似文献   

9.
The interaction of melittin, a polypeptide consisting of 26 amino acid residues, with dimyristoyl phosphatidylcholine bilayers was investigated by vibrational Raman spectroscopy. Spectral peak height intensity ratios, involving vibrational transitions in both the 3000 cm?1 acyl chain methylene carbon-hydrogen stretching mode region and the 1100 cm?1 acyl chain carbon-carbon skeletal stretching mode interval, served as temperature profile indices for monitoring the bilayer order-disorder processes. For a lipid : melittin molar ratio of 14 : 1 two order-disorder transitions were observed. In comparison to a gel to liquid crystalline phase transition of 22.5°C for the pure lipid, the lower transition, exhibiting a 2°C width, is centered at 17°C and is associated with a depression of the main lipid phase transition of dimyristoyl phosphatidylcholine. The second thermal transition, displaying a 7°C interval, occurs at approx. 29°C and is associated with the melting behavior of approximately seven immobilized boundary lipids which surround the inserted hydrophobic segment of the polypeptide. For a lipid : melittin molar ratio of 10 : 1 two thermal transitions are also observed at 11 and 30°C. As before, they represent, respectively, the main gel to liquid crystalline phase transition and the melting behavior of approximately four boundary lipids attached to melittin. From these data alternative schemes are suggested for disposing the immobilized lipids around the hydrophobic portion of the polypeptide within the bilayer.  相似文献   

10.
The interaction of the probe diS-C3-(5) with dipalmitoylphosphatidylcholine (DPPC) liposomes has been studied using fluorescence and differential scanning calorimetry (DSC). The partition coefficients (K) of the probe for the lipid and the aqueous phase (in terms of molar part units) were (1.20 +/- 0.4) X 10(6) at 45 degrees C and (0.50 +/- 0.07) X 10(6) at 23 and 36 degrees C. In terms of volume concentration units, these values correspond to Kp = (2.88 +/- 0.10) X 10(4) and Kp = (1.20 +/- 0.17) X 10(4), respectively. DSC thermograms were practically identical both for large unilamellar and multilamellar liposomes. The main transition peak remained practically unchanged over the entire range of the probe concentrations used. The pretransition could be observed up to maximal probe concentrations applied and it widened and shifted from 35.4 degrees C in pure DPPC to approximately 32 degrees C at a probe/lipid ratio of 0.027. These results suggest that in both quasicrystalline and liquid crystalline lipid bilayers the probe molecules are included in "defects" between structurally ordered microregions (microdomains or clusters). The dependence of the fluorescence response on the transmembrane potential in a suspension of unilamellar DPPC vesicles suggest that the equilibrium thermodynamic model is valid for liquid crystalline bilayers.  相似文献   

11.
Proton magnetic resonance spectra of ubiquinone-10 and ubiquinone-10 dispersed with dipalmitoylglycerophosphocholine or egg phosphatidylcholine in aqueous medium have been obtained. The dispersions are in the form of multilamellar liposomes as judged by 31P-NMR spectra and the thermal history of the samples have ensured that ubiquinone not incorporated into the phospholipid structure only gives rise to a broad-line NMR proton spectrum. A high-resolution proton spectrum of ubiquinone is observed with upfield shifts of the O-methyl protons of the benzoquinone rings, indicating close proximity of the molecules but with an arrangement different from the pure liquid ubiquinone. Spectra obtained in the presence of the lanthanide shift reagents, dysprosium fluorooctanedionate and Dy(NO3)3, which have a preferred location in the hydrophobic and hydrophilic domains, respectively, of ubiquinone/phospholipid codispersions, are consistent with the partitioning of ubiquinone into a hydrophobic phospholipid environment remote from the aqueous phase. The type of arrangements of ubiquinone that could be accommodated within bilayers of phospholipid are discussed.  相似文献   

12.
The temperature-composition phase diagram of monopentadecenoin, a monoacylglycerol with a cis monounsaturated fatty acid 15 carbon atoms long (C15:1c10) in water was constructed using x-ray diffraction. Low- and wide-angle diffraction patterns were collected from samples of fixed hydration as a function of temperature in the heating direction on x-ray-sensitive film. The temperature and hydration ranges investigated were 0-104 degrees C and 0-60% (w/w) water, respectively. The phases identified in the system include the lamellar crystalline phase, the lamellar liquid crystalline phase, the fluid isotropic phase, and two inverted cubic phases belonging to space groups la3d (Q230) and Pn3m (Q244). Particular attention has been devoted to the issues of phase equilibrium, phase boundary verification, and structure characterization. The phase diagrams of monopentadecenoin, monomyristolein (C14:1c9), and monoolein (C18:1c9) are compared, and the impact of molecular structure on mesophase stability and structure is discussed.  相似文献   

13.
Liposomes containing phospholipids with covalently attached poly(ethylene glycol) (PEG-lipids) are being developed for in vivo drug delivery. In this paper we determine the structure and phase behavior of fully hydrated distearoylphosphatidylcholine (DSPC) suspensions containing PEG-lipids composed of distearoylphosphatidylethanolamine with attached PEGs of molecular weights ranging from 350 to 5000. For DSPC:PEG-lipid suspensions containing 0-60 mol % PEG-lipid, differential scanning calorimetry shows main endothermic transitions ranging from 55 to 64 degrees C, depending on the size of the PEG and concentration of PEG-lipid. The enthalpy of this main transition remains constant for all PEG-350 concentrations but decreases with increasing amounts of PEG-750, PEG-2000, or PEG-5000, ultimately disappearing at PEG-lipid concentrations greater than about 60 mol %. Low-angle and wide-angle x-ray diffraction show that tilted gel (L beta') phase bilayers are formed for all PEG-lipid molecular weights at concentrations of about 10 mol % or less, with the distance between bilayers depending on PEG molecular weight and PEG-lipid concentration. At PEG-lipid concentrations greater than 10 mol %, the lipid structure depends on the size of the PEG moiety. X-ray diffraction analysis shows that untilted interdigitated (L beta I) gel phase bilayers form with the incorporation of 40-100 mol % PEG-350 or 20-70 mol % PEG-750, and untilted gel (L beta) phase bilayers are formed in the presence of about 20-60 mol % PEG-2000 and PEG-5000. Light microscopy, turbidity measurements, x-ray diffraction, and 1H-NMR indicate that a pure micellar phase forms in the presence of greater than about 60% PEG-750, PEG-2000, or PEG-5000.  相似文献   

14.
X-ray diffraction measurements show that the β′-crystalline to liquid crystalline transitions of dipalmitoyl lecithin-water mixtures take place over temperature ranges of several degrees, and by the gradual disappearance of one crystalline state accompanied by the gradual appearance of the other. The narrowness of the low angle X-ray lines indicates that the two crystalline states exist in separate phases. At a given temperature, the liquid crystalline phase has the greater water content. The bilayers of the liquid crystalline phase are thinner than those of the β′-crystalline phase while the distances between bilayers are essentially equal in the two phases.  相似文献   

15.
The membranes of the thermoacidophilic archaeon Sulfolobus metallicus exhibit an oxygen consumption activity of 0.5 nmol O(2) min(-1) mg(-1), which is insensitive to rotenone, suggesting the presence of a type-II NADH dehydrogenase. Following this observation, the enzyme was purified from solubilised membranes and characterised. The pure protein is a monomer with an apparent molecular mass of 49 kDa, having a high N-terminal amino acid sequence similarity towards other prokaryotic enzymes of the same type. It contains a covalently attached flavin, which was identified as being FMN by 31P-NMR spectroscopy, a novelty among type-II NADH dehydrogenases. Metal analysis showed the absence of iron, indicating that no FeS clusters are present in the protein. The average reduction potential of the FMN group was determined to be +160 mV, at 25 degrees C and pH 6.5, by redox titrations monitored by visible spectroscopy. Catalytically, the enzyme is a NADH:quinone oxidoreductase, as it is capable of transferring electrons from NADH to several quinones, including ubiquinone-1, ubiquinone-2 and caldariella quinone. Maximal turnover rates of 195 micromol NADH oxidized min(-1) mg(-1) at 60 degrees C were obtained using ubiquinone-2 as electron acceptor, after enzyme dilution and incubation with phospholipids.  相似文献   

16.
Sherman MB  Orlova EV  Decker GL  Chiu W  Pownall HJ 《Biochemistry》2003,42(50):14988-14993
Low-density lipoprotein (LDL) particles from normolipidemic individuals contain a cholesteryl ester-rich core that undergoes a thermal transition from a liquid crystalline to an isotropic liquid phase between 20 and 35 degrees C. LDL from hypertriglyceridemic patients or prepared in vitro by the exchange of very low-density lipoprotein for LDL cholesteryl esters is triglyceride-rich, does not have a thermal transition above 0 degrees C, and exhibits impaired binding to the LDL receptor on normal human skin fibroblasts. Cryoelectron microscopy of LDL quick-frozen from 10 (core-frozen) and 40 degrees C (core-melted) revealed ellipsoidal particles with internal striations and round particles devoid of striations, respectively. Cryoelectron microscopy of triglyceride-rich LDL prepared in vitro revealed particles similar to the core-melted normolipidemic LDL, i.e., round particles without striations. These data suggest that the LDL core in the liquid crystalline phase is characterized by the appearance of striations, whereas LDL with a core that is an isotropic liquid lacks striations. It is suggested that freezing the LDL core into a liquid crystalline phase imposes structural constraints that force LDL from a sphere without partitions to an ellipsoid with partitions. We further suggest that the striation-defined lamellae are a structural feature of a liquid crystalline neutral lipid core that is a determinant of normal binding to the LDL receptor and that conversion of the neutral lipid core of LDL to the isotropic liquid phase via an increase in the temperature or via the addition of triglyceride partially ablates the receptor binding determinants on the LDL surface. This effect is likely achieved through changes in the conformation of apo-B-100. These data suggest that the physical state of the LDL core determines particle shape, surface structure, and metabolic fate.  相似文献   

17.
The spontaneous reconstitution of lipid-protein complexes was examined by mixing bacteriorhodopsin or UDP-glucuronosyltransferase with preformed, unilamellar bilayers of pure dimyristoylphosphatidylcholine. Spontaneous insertion of these proteins into vesicles of dimyristoylphosphatidylcholine was facilitated by resonicating the vesicles at 4 degrees C. The property of resonicated vesicles that led to spontaneous reconstitution could be annealed by melting the bilayers, which slowed down reconstitution. The overall process of reconstitution consisted, however, of two steps. There was an initial insertion of proteins into a small portion of vesicles followed by subsequent fusion between protein-free vesicles and vesicles containing lipid-protein complexes. The first step appeared to proceed rapidly in all vesicles in a gel phase, whether or not they were resonicated or whether or not resonicated vesicles were annealed. The rate of the second step was sensitive to these treatments. The membrane proteins also inserted into preformed vesicles in a liquid crystalline phase, but this step was slower than for vesicles in a gel phase. Fusion between protein-free and protein-containing vesicles in a liquid crystalline phase was extremely slow. The data show that the spontaneous insertion of pure membrane proteins into preformed vesicles can be a facile event and that the overall reconstitution of membrane proteins into preformed unilamellar vesicles may be simpler to achieve than has been appreciated.  相似文献   

18.
M F Brown  J Seelig 《Biochemistry》1978,17(2):381-384
The structural changes in the polar head group region of unsonicated bilayer membranes of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine and 1,2-dipalmitoyl-sn-glycero-3-phosphoethanolamine produced by addition of cholesterol have been determined using deuterium and phosphorus-31 NMR. Incorportion of up to 50 mol percent cholesterol produces little change in the phosphorus-31 chemical shielding anisotropies, compared with the values in pure bilayers above the phase transition temperatures, while some of the deuterium quadrupole splittings are reduced by almost a factor of two. Adjustment of the head group torsion angles by only a few degrees accounts for the observed spectral changes. Addition of cholesterol therefore has opposite effects on the hydrocarbon and polar regions of membranes: although cholesterol makes the hydrocarbon region gel-like, with an increased probability of trans conformations, the conformation of the polar head groups is very similar to that found in the liquid crystalline phase of pure phospholipid bilayers.  相似文献   

19.
(31)P solid-state NMR spectroscopy has been used to investigate the macroscopic phase behavior of phospholipid bilayers in the presence of increasing amounts of magainin antibiotic peptides. Addition of >1 mol% magainin 2 to gel-phase DMPC or liquid crystalline POPC membranes respectively, results in (31)P NMR spectra that are characterized by the coexistence of isotropic signals and line shapes typical for phospholipid bilayers. The isotropic signal intensity is a function of temperature and peptide concentration. At peptide concentrations >4 mol% of the resulting phospholipid (31)P NMR spectra are characteristic of magnetically oriented POPC bilayers suggesting the formation of small disk-like micelles or perforated sheets. In contrast, addition of magainin to acidic phospholipids results in homogenous bilayer-type (31)P NMR spectra with reduced chemical shift anisotropies. The results presented are in good agreement with the interfacial insertion of magainin helices with an alignment parallel to the surface of the phospholipid bilayers. The resulting curvature strain results in detergent-like properties of the amphipathic helical peptides.  相似文献   

20.
Monolayers of dipalmitoylphosphatidylcholine (DPPC), dioleoylphosphatidylcholine (DOPC), and some mixtures of these lipids were investigated using an epifluorescence microscopic surface balance. Monolayers were visualized at 23 +/- 1 degree C through the fluorescence of 1 mol% of two different fluorescent probes, 1-palmitoyl-2-(12-[(7-nitro-2-1,3-benzoxadizole-4- yl)amino]dodecanoyl)phosphatidylcholine (NBD-PC), which partitions into the liquid expanded (LE) or disordered lipid phase and 3,3'-dioctadecyloxacarbocyanine perchlorate (DiO-C18), which preferentially associates with the liquid condensed (LC) phase or lipid with ordered chains. LC domains were observed in pure DPPC monolayers at relatively low surface pressures (pi), and these domains grew with increasing surface pressure. Only liquid expanded phase was observed in pure DOPC monolayers up to the point of monolayer collapse. In monolayers containing 29:70:1, 49:50:1, and 69:30:1 (mol/mol/mol) of DPPC:DOPC:probe the domains of LC phase were smaller than those seen in DPPC monolayers at equivalent surface pressures. Quantitative analysis of the visual fields shown by the mixed monolayers showed a distribution of sizes of condensed domains at any given pi. At pi = 30 mN m-1, liquid-expanded, or fluid, regions occupied more than 70% of the total monolayer area in all three mixtures studied, whereas DPPC monolayers were more than 75% condensed or solid at that pressure. For monolayers of DPPC:DOPC:NBD-PC 49:50:1 and 69:30:1 the average domain size and the percentage of the total area covered with LC, or rigid, areas increased to a maximum at pi around 35 mN m-1 followed by a decrease at higher pi.(ABSTRACT TRUNCATED AT 250 WORDS)  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