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 46 毫秒
1.
Photovoltaic conversion using zinc chlorin-e6 (ZnChl-e6), which is zinc chlorophyll-a derivative, and fatty acid (myristic acid or cholic acid) co-adsorbed nanocrystalline TiO2 layer onto ITO glass (OTE) electrode is developed. The maximum peaks of photocurrent action spectrum of the ZnChl-e6 adsorbed TiO2 layer onto OTE (ZnChl-e6/TiO2) are 400, 660 and 800 nm, respectively. Especially the IPCE value at 800 nm (7.5%) is larger than that of 660 nm (6.9%). This result indicates that ZnChl-e6 molecules is aggregated or formed dimer on a nanocrystalline TiO2 layer onto OTE and the absorption band is shifted to near IR region. The photocurrent action spectrum of ZnChl-e6 and cholic acid adsorbed TiO2 layer onto OTE (ZnChl-e6-Cho/TiO2 is similar to that of the UV-vis absorption spectrum in methanol solution, and IPCE values at 400 and 660 nm (8.1%) increase and the IPCE value at 800 nm (4.1%) decreases, indicating that the aggregation of ZnChl-e6 molecules on the TiO2 is suppressed by cholic acid. By using ZnChl-e6-Cho/TiO2, the short-circuit photocurrent density and open-circuit photovoltage also increase compared with that of ZnChl-e6 adsorbed nanocrystalline TiO2 electrode.  相似文献   

2.
A bio-photovoltaic conversion device based on dye-sensitised solar cell (DSSC) using the visible light sensitisation of chlorine-e6 (Chl-e6) derived from chlorophyll from Spirulina adsorbed on a nanocrystalline TiO2 film was developed. Form fluorescence spectrum of Chl-e6 adsorbed on a nanocrystalline TiO2 film, the emission of Chl-e6 was effectively quenched by TiO2 nanocrystalline indicating that the effective electron injection from the excited singlet state of Chl-e6 into the conduction band of TiO2 particles occurred. The short-circuit photocurrent density (Isc). the open-circuit photovoltage (Voc). and the fill factor (FF) of solar cell using Chl-e6 adsorbed on a nanocrystalline TiO2 film electrode were estimated to be 0.305 +/- 0.012 mA cm(-2), 426 +/- 10 mV, and 45.0%, respectively. IPCE values were reached a maximum around the wavelength of absorption maximum (7.40% at 400 nm; 1.44% at 514 nm and 2.91% at 670 nm), indicating that the DSSC using visible light sensitisation of nanocrystalline TiO2 film by Chl-e6 was developed.  相似文献   

3.
A decline in CFU of gram-positive and gram-negative bacteria on the surface of UV illuminated TiO2 films (wavelength of 380 nm) is shown. A 29, 45, and 47% decrease in bacterial viability of Staphylococcus aureus, Staphylococcus epidermidis, and Escherichia coli, respectively, was seen following a 12-min exposition. It was first discovered that the reuse of TiO2 films to test a bacterial suspension for viability removes UV-induced bactericidal activity. However, annealing of TiO2 at a temperature above 400 degrees C restores the photoinduced bactericidal activity to its initial state.  相似文献   

4.
Thermal stability of myosin rod from various species   总被引:1,自引:0,他引:1  
The radius of gyration and fraction helix as a function of temperature have been determined for myosin rod from four different species: rabbit, frog, scallop, and antarctic fish. Measurements from sodium dodecyl sulfate gel electrophoresis indicate that all particles have the same molecular weight (approximately 130K). All fragments are nearly 100% alpha-helical at low temperatures (0-5 degrees C). The melting profiles for each are qualitatively similar in shape, but their midpoints are shifted along the temperature axis in the following order: antarctic fish (Tm = 33 degrees C), scallop (Tm = 39 degrees C), frog (Tm = 45 degrees C), and rabbit (Tm = 49 degrees C). Corresponding radius of gyration vs temperature profiles for each species are shifted to lower temperatures (approximately 5-8 degrees C) with respect to the optical rotation melting curves. From plots of radius of gyration vs fraction helix, we find a marked drop in the radius of gyration (from 43 to approximately 34 nm) with less than a 5% decrease in fraction helix for rabbit, frog, and antarctic fish rods, whereas the radius of gyration of scallop rod never exceeds 34 nm. Results indicate hinging of the myosin rod of each species. The thermal stabilities of the myosin rods shift in parallel with the working temperature of their respective muscles.  相似文献   

5.
Luminescence emission, polarization, and excitation spectra of polyadenylic acid (poly(A)) have been studied in room-temperature aqueous solution (pH 8). The temperature dependence of the luminescence of poly(A) at two different excitation wavelengths in the range 10-70 degrees C has also been studied and compared with that of adenosine 5'-monophosphate (AMP). It has been found that the luminescence excitation spectrum and the polarization curve of poly(A) solution reveal a low-intensity electronic transition at about 320 nm which is red-shifted by approximately 0.9 eV from the maximum of the absorption spectrum at 260 nm. A model of two luminescent stacked forms is suggested. The difference in the ground state levels of these two stacked forms obtained from the temperature dependencies of the emissions is insignificant ( approximately 1 kcal/mol). This means a lowering of the excited state of the stacked form with the 320 nm/420 nm absorption/emission bands by approximately 0.9 eV as compared to the main form with the 260 nm/400 nm absorption/emission bands. The low-lying excited states suggest the stronger electronic coupling of the bases in a certain stacked form. It is proposed that such clusters of the stacked bases could provide the wire-like conductivity in the short segments of DNA.  相似文献   

6.
The structural, surface morphological, optical absorption and emission features of Y2O3:Ce (0%–5%) were studied. The samples had a body-centred cubic crystal structure. The undoped sample had a crystallite size of 29.03 nm, and it varied after doping with Ce. The grain size of the samples varied from 23.00 to 50.78 nm. All the samples exhibited a strong absorption band at 206 nm due to F-centre absorption and absorption involving the delocalised bands. In addition, the doped samples exhibited a secondary band at ~250 nm due to 4f → 5d transitions of Ce3+ ions. The optical bandgap of the undoped sample was found to be ~5.37 eV, and it decreased to 5.20 eV with an increase in Ce concentration to 5%. The undoped sample under 350-nm excitation exhibited a broad photoluminescence (PL) emission band with the maxima at 406 nm and a secondary band at 463 nm. In contrast, multiple PL peaks were centred at ~397, 436, 466, 488 and 563 nm in all the doped samples. The average lifetime of the emission band at 406 nm was 1.05 ns and that of the emission band at ~466 nm was 1.63 ns. The material has potential for solid-state lighting applications.  相似文献   

7.
The Caulobacter polar surface structures (flagella, pili, and the deoxyribonucleic acid phage phiCbK receptors), which are expressed at proximal sites of swarmer cells in a coordinate manner (Shapiro, Annu. Rev. Microbiol., 30:377-407, 1976) could be blocked by a single mutation. The mutant C. crescentus CB13 ple-801 did not form these surface structures when grown at 35 degrees C. Upon shift down to 25 degrees C, the mutant cells initiated the formation of the surface structures. When mitomycin C was added to the mutant culture upon shift down from 35 to 25 degrees C, phiCbK receptor formation was inhibited to a minimal level. Rifampin and chloramphenicol completely inhibited phiCbK receptor formation when added to the mutant culture upon shift down. Deoxyribonucleic acid as well as ribonucleic acid and protein synthesis seem to be required for the formation of phiCbK receptors. Penicillin V also inhibited phiCbK receptor formation, indicating the involvement of cell wall synthesis. When the mutant CB13 ple-801 cells were shifted down briefly from 35 to 25 degrees C and then shifted up to 35 degrees C, flagella and phiCbK receptors were formed even at 35 degrees C to different extents depending on how long the cells were incubated at 25 degrees C. This formation of the surface structures at 35 degrees C was inhibited by rifampin. From these results, it appears that translation, assembly, or localization processes for the formation of the surface structures are not temperature sensitive at 35 degrees C in the pleiotropic mutant CB13 ple-801. The syntheses of deoxyribonucleic acid and the cell wall do not appear to be temperature sensitive either, since the mutant grows normally at 35 degrees C. It is suggested that there exists a regulatory step that commits the cells to initiate the synthesis of requisite ribonucleic acid for the formation of the polar surface structures.  相似文献   

8.
The purpose of this research was to improve the stability of carbamazepine (CBZ) bulk powder under high humidity by surface modification. The surface-modified anhydrates of CBZ were obtained in a specially designed surface modification apparatus at 60°C via the adsorption of n-butanol, and powder x-ray diffraction, Fourier-Transformed Infrared spectra, and differential scanning calorimetry were used to determine the crystalline characteristics of the samples. The hydration process of intact and surface-modified CBZ anhydrate at 97% relative humidity (RH) and 40±1°C was automatically monitored by using isothermal microcalorimetry (IMC). The dissolution test for surface-modified samples (20 mg) was performed in 900 mL of distilled water at 37±0.5°C with stirring by a paddle at 100 rpm as in the Japanese Pharmacopoeia XIII. The heat flow profiles of hydration of intact and surface-modified CBZ anhydrates at 97% RH by using IMC profiles showed a maximum peak at around 10 hours and 45 hours after 0 and 10 hours of induction, respectively. The result indicated that hydration of CBZ anhydrate was completely inhibited at the initial stage by surface modification of n-butanol and thereafter transformed into dihydrate. The hydration of surface-modified samples followed a 2-dimensional phase boundary process with an induction period (IP). The IP of intact and surface-modified samples decreased with increase of the reaction temperature, and the hydration rate constant (k) increased with increase of the temperature. The crystal growth rate constants of nuclei of the intact sample were significantly larger than the surface-modified samples at each temperature. The activation energy (E) of nuclei formation and crystal growth process for hydration of surface-modified CBZ anhydrate were evaluated to be 20.1 and 32.5 kJ/mol, respectively, from Arrhenius plots, but the Es of intact anhydrate were 56.3 and 26.8 kJ/mol, respectively. The dissolution profiles showed that the surface-modified sample dissolved faster than the intact sample at the initial stage. The dissolution kinetics were analyzed based on the Hixon-Crowell equation, and the dissolution rate constants for intact and surface-modified anhydrates were found to be 0.0102±0.008 mg1/3 min−1 and 0.1442±0.0482 mg1/3·min−1. The surface-modified anhydrate powders were more stable than the nonmodified samples under high humidity and showed resistance against moisture. However, surface modification induced rapid dissolution in water compared to the control.  相似文献   

9.
The effect of extraction of weakly bound Ca2+ by low-pH treatment on the O2-evolving apparatus was studied by use of low-temperature electron paramagnetic resonance (EPR) and X-ray absorption spectroscopy. In low-pH-treated PSII membranes, an S2 EPR multiline signal with modified line shape was induced by illumination at 0 degrees C, but its signal amplitude decreased upon lowering the excitation temperature with concomitant oxidation of cytochrome (cyt) b-559 in place of Mn. The half-inhibition temperature for formation of the modified multiline signal was found at -33 degrees C, which was much higher than that for formation of the normal S2 state in untreated control membranes. Signal IIf was normally induced down to -30 degrees C, but its dependence on excitation temperature was different from that for modified S2. This was interpreted as indicating that the low-temperature blockage of modified S2 formation is due to the incapability of electron abstraction from the Mn cluster. The Mn K-edge of X-ray absorption near-edge structure (XANES) spectrum shifted to lower energy by 0.8 eV after low-pH treatment, but the shift was reversed by addition of Ca2+. Upon illumination at 0 degrees C of treated membranes, the K-edge energy was up-shifted by 0.8 eV, but was not upon illumination at 210 K. These results were interpreted as indicating that extraction of weakly bound Ca2+ by low-pH treatment gives rise to structural and functional modulations of the Mn cluster.  相似文献   

10.
Displacements of single one-headed myosin molecules in a sparse myosin-rod cofilament were measured from bead displacements at various angles relative to an actin filament by dual optical trapping nanometry. The sparse myosin-rod cofilaments, 5-8 micron long, were synthesized by slowly mixing one-headed myosin prepared by papain digestion with myosin rods at molar ratios of 1:400 to 1:1500, so that one to four one-headed myosin molecules were on average scattered along the cofilament. The bead displacement was approximately 10 nm at low loads ( approximately 0.5 pN) and at angles of 5-10 degrees between the actin and myosin filaments (near physiologically correct orientation). The bead displacement decreased with an increase in the angle. The bead displacement at nearly 90 degrees was approximately 0 nm. When the angle was increased to approximately 150 degrees-170 degrees, the bead displacements increased to 5 nm. A native two-headed myosin showed similar size and orientation dependence of bead displacements as a one-headed myosin.  相似文献   

11.
Electron inelastic cross sections and stopping powers for liquid water over the 0.1-10 keV range are presented based on a recently developed dielectric response model for liquid water (D. Emfietzoglou, F. Cucinotta and H. Nikjoo, Radiat. Res. 164, 202-211, 2005) that is consistent with the experimental data over the whole energy-momentum plane. Both exchange and second-order Born corrections are included in a material-specific way using the dielectric functions of liquid water. The numerical results are fitted by simple analytic functions to facilitate their further use. Compared to previous studies, differential cross sections are shifted toward smaller energy losses resulting in smaller inelastic and stopping cross sections with differences reaching, on average, the approximately 20% and approximately 50% level, respectively. Contrary to higher energies, it is shown that the dispersion model for the momentum dependence of the dielectric functions (Bethe ridge) is as important as the optical model used. Within the accuracy of the experimental data (a few percent) upon which our dielectric model is based, the calculations are "exact" to first order, while the uncertainty of the results beyond first order is estimated at the 5-10% level. The present work overcomes the limitations of Bethe's theory at low energies by a self-consistent account of inner-shell effects and may serve to extend the ICRU electron stopping power database for liquid water down to 100 eV with a level of uncertainty similar to that for the higher-energy values.  相似文献   

12.
The efficiency of PAMAM dendrimer-mediated DNA transfer can be improved by the addition of substituted beta-cyclodextrins (beta-CDs) as formulation excipients. In vitro CAT expression increased approximately 200-fold when dendrimer/DNA/beta-CD formulations were applied on the surface of collagen membranes. The inclusion of beta-CD into the formulations resulted in particles that were smaller and more evenly distributed on the surface of the solid support. The average size of the complex formed at 50 microg/ml and at charge ratio of 1 decreased from 156 nm to 5.8 nm and 21.2 nm in 0.025-0.1% w/vol beta-CDs. Sulfonated beta-CDs bind to dendrimer and in the increased concentration may displace DNA in the dendrimer/DNA complex. High concentrations of amphoteric beta-CD do not dissociate dendrimer/DNA complexes; however, they may decrease their ability to transfect cells. At the optimized formulations the surface-modified beta-CDs may enhance solid support-based transfection in vitro, through modification of dendrimer/DNA complex composition and improved surface distribution.  相似文献   

13.
Carrier dynamics in methylammonium lead halide (CH3NH3PbI3–xClx) perovskite thin films, of differing crystal morphology, are examined as functions of temperature and excitation wavelength. At room temperature, long‐lived (>nanosecond) transient absorption signals indicate negligible carrier trapping. However, in measurements of ultrafast photoluminescence excited at 400 nm, a heretofore unexplained, large amplitude (50%–60%), 45 ps decay process is observed. This feature persists for temperatures down to the orthorhombic phase transition. Varying pump photon energy reveals that the fast, band‐edge photoluminescence (PL) decay only appears for excitation ≥2.38 eV (520 nm), with larger amplitudes for higher pump energies. Lower photon‐energy excitation yields slow dynamics consistent with negligible carrier trapping. Further, sub‐bandgap two‐photon pumping yields identical PL dynamics as direct absorption, signifying sensitivity to the total deposited energy and insensitivity to interfacial effects. Together with first principles electronic structure and ab initio molecular dynamics calculations, the results suggest the fast PL decay stems from excitation of high energy phonon modes associated with the organic sub‐lattice that temporarily enhance wavefunction overlap within the inorganic component owing to atomic displacement, thereby transiently changing the PL radiative rate during thermalization. Hence, the fast PL decay relates a characteristic organic‐to‐inorganic sub‐lattice equilibration timescale at optoelectronic‐relevant excitation energies.  相似文献   

14.
The Escherichia coli DEAD-box protein A (DbpA) is an RNA helicase that utilizes the energy from ATP binding and hydrolysis to facilitate structural rearrangements of rRNA. We have used the fluorescent nucleotide analogues, mantADP and mantATP, to measure the equilibrium binding affinity and kinetic mechanism of nucleotide binding to DbpA in the absence of RNA. Binding generates an enhancement in mant-nucleotide fluorescence and a corresponding reduction in intrinsic DbpA fluorescence, consistent with fluorescence resonance energy transfer (FRET) from DbpA tryptophan(s) to bound nucleotides. Fluorescent modification does not significantly interfere with the affinities and kinetics of nucleotide binding. Different energy transfer efficiencies between DbpA-mantATP and DbpA-mantADP complexes suggest that DbpA adopts nucleotide-dependent conformations. ADP binds (K(d) approximately 50 microM at 22 degrees C) 4-7 times more tightly than ATP (K(d) approximately 400 microM at 22 degrees C). Both nucleotides bind with relatively temperature-independent association rate constants (approximately 1-3 microM(-1) s(-1)) that are much lower than predicted for a diffusion-limited reaction. Differences in the binding affinities are dictated primarily by the dissociation rate constants. ADP binding occurs with a positive change in the heat capacity, presumably reflecting a nucleotide-induced conformational rearrangement of DbpA. At low temperatures (<22 degrees C), the binding free energies are dominated by favorable enthalpic and unfavorable entropic contributions. At physiological temperatures (>22 degrees C), ADP binding occurs with positive entropy changes. We favor a mechanism in which ADP binding increases the conformational flexibility and dynamics of DbpA.  相似文献   

15.
Elevated partial pressures of atmospheric carbon dioxide, similar to numerous causes of plant stress, may alter leaf pigmentation and structure and thus would be expected to alter leaf optical properties. Hypotheses that elevated CO(2) pressure and air temperature would alter leaf optical properties were tested for sugar maple (Acer saccharum) in the middle of its fourth growing season under treatment. The saplings had been growing since 1994 in open-top chambers and partial shade at Oak Ridge, Tennessee under the following treatments: (1) ambient CO(2) pressure and air temperature (control); (2) CO(2) pressure approximately 30 Pa above ambient; (3) air temperatures 3 degrees C above ambient; and (4) elevated CO(2) and air temperature. Under elevated CO(2) or temperature, spectral reflectance, transmittance and absorptance in the visible spectrum (400-720 nm) tended to change in patterns that generally are associated with chlorosis, with maximum differences from the control near 700 nm. However, these changes were not significant at P=0.05. Although reflectance, transmittance and absorptance at 700 nm correlated strongly with leaf chlorophyll concentration, variability in chlorophyll concentration was greater within than among treatments. The lack of treatment effects on pigmentation explained the non-significant change in optical properties in the visible spectrum. Optical properties in the near-infrared (721-850 nm) were similarly unresponsive to treatment with the exception of an increased absorptance throughout the 739-850 nm range in leaves that developed under elevated air temperature alone. This response might have resulted from effects of air temperature on leaf internal structure.  相似文献   

16.
The question of hinging in myosin rod from rabbit skeletal muscle has been reexamined. Elastic light scattering and optical rotation have been used to measure the radius of gyration and fraction helix, respectively, as a function of temperature for myosin rod, light meromyosin (LMM), and long subfragment 2 (long S-2). The radius of gyration vs temperature profile of myosin rod is shifted with respect to the optical rotation melting curve by about -5 degrees C. Similar studies on both LMM and long S-2 show virtually superimposable profiles. To correlate changes in the secondary structure with the overall conformation, plots of radius of gyration vs fraction helix are presented for each myosin subfragment. Myosin rod exhibits a marked decrease in the radius of gyration from 43 nm to approximately 35 nm, while the fraction helix remains at nearly 100%. LMM and long S-2 did not show this behavior. Rather, a decrease in the radius of gyration of these fragments occurred with comparable changes in fraction helix. These results are interpreted in terms of hinging of the myosin rod within the LMM/S-2 junction.  相似文献   

17.
Highly dispersible TiO2 nanocrystals (approximately 6 nm) were prepared by an alcohothermal method. A strong and stable photoluminescence emission with a maximum at 450 nm was observed in the original TiO2 nanocrystals colloid. Compared with the emission from quinine sulphate in 0.05 mol/L sulphuric acid, the emission quantum yield of TiO2 nanocrystals was determined to be about 0.20, which was much higher than the values (0.002 and 0.001) reported in previous studies. The fluorescence micrograph of TiO2 nanocrystals encapsulated in lipsomes shows that TiO2 nanocrystals prepared in such a way have potential application as a fluorescence probe in biological imaging. Research on the luminescence mechanism indicates that the surface state and extent of crystallization of the TiO2 nanocrystals are crucial factors in the high photoluminescence quantum yield obtained in this study.  相似文献   

18.
Prothrombin denaturation was examined in the presence of Na2EDTA, 5mM CaCl2, and CaCl2 plus membranes containing 1-palmitoyl-2-oleoyl-3-sn-phosphatidylcholine (POPC) in combination with either bovine brain phosphatidylserine (PS) or 1,2-dioleoyl-phosphatidylglycerol (DOPG). Heating denaturation of prothrombin produced thermograms showing two peaks, a minor one at approximately 59 degrees C previously reported to correspond to denaturation of the fragment 1 region (Ploplis, V. A., D. K. Strickland, and F. J. Castellino 1981. Biochemistry. 20:15-21), and a main one at approximately 57-58 degrees C, reportedly due to denaturation of the rest of the molecule (prethrombin 1). The main peak was insensitive to the presence of 5mM Ca2+ whereas the minor peak was shifted to higher temperature (Tm approximately 65 degrees C) by Ca2+. Sufficient concentrations of POPC/bovPS (75/25) large unilamellar vesicles to guarantee binding of 95% of prothrombin resulted in an enthalpy loss in the main endotherm and a comparable enthalpy gain in the minor endotherm accompanying an upward shift in peak temperature (Tm approximately 73 degrees C). Peak deconvolution analysis on the prothrombin denaturation profile and comparison with isolated prothrombin fragment 1 denaturation endotherms suggested that the change caused by POPC/PS vesicles reflected a shift of a portion of the enthalpy of the prethrombin 1 domain to higher temperature (Tm approximately 77 degrees C). The enthalpy associated with this high-temperature endotherm increased in proportion to the surface concentration of PS. By contrast, POPC/DOPG (50/50) membranes shifted the prethrombin 1 peak by 4 degrees C to a lower temperature and the fragment 1 peak by 5 degrees C to a higher temperature. The data lead to a hypothesis that the fragment 1 and prethrombin 1 domains of prothrombin do not denature quite independently and that binding of prothrombin to acidic-lipid membranes disrupts the interaction between these domains. It is further hypothesized that PS containing membranes exert the additional specific effect of decoupling the denaturation of two subdomains of the prethrombin 1 domain of prothrombin.  相似文献   

19.
Canes from Arundo donax, a herbaceous rapid-growing plant, were used as precursor for activated carbon preparation by phosphoric acid activation under a self-generated atmosphere. The influence of the carbonization temperature in the range 400-550 degrees C and of the weight ratio phosphoric acid to precursor (R = 1.5-2.5) on the developed porous structure of the resulting carbons was studied for 1 h of carbonization time. Surface properties of the activated carbons were dependent on a combined effect of the conditions employed. Carbons developed either with R = 1.5 over the range 400-500 degrees C, or with R = 2 at 500 degrees C exhibited surface areas of around 1100 m2/g, the latter conditions promoting a larger pore volume and enhanced mesoporous character. For both ratios, temperature above 500 degrees C led to reduction in porosity development. A similar effect was found for the highest ratio (R = 2.5) and 500 degrees C. The influence of carrying out the carbonization either for times shorter than 1 h or under flowing N2 was also examined at selected conditions (R = 2, 500 degrees C). Shorter times induced increase in the surface area (approximately 1300 m2/g), yielding carbons with smaller mean pore radius. Activated carbons obtained under flowing N2 possessed predominant microporous structures and larger ash contents than the samples derived in the self-generated atmosphere.  相似文献   

20.
G Cevc  A Watts  D Marsh 《Biochemistry》1981,20(17):4955-4965
The dependence of the gel-to-fluid phase transition temperature of dimyristoyl- and dipalmitoylphosphatidylserine bilayers on pH, NaCl concentration, and degree of hydration has been studied with differential scanning calorimetry and with spin-labels. On protonation of the carboxyl group (pK2app = 5.5), the transition temperature increases from 36 to 44 degrees C in the fully hydrated state of dimyristoylphosphatidylserine (from 54 to 62 degrees C for dipalmitoylphosphatidylserine), at ionic strength J = 0.1. In addition, at least two less hydrated states, differing progressively by 1 H2O/PS, are observed at low pH with transition temperatures of 48 and 52 degrees C for dimyristoyl- and 65 and 68.5 degrees C for dipalmitoylphosphatidylserine. On deprotonation of the amino group (pK3app = 11.55) the transition temperature decreases to approximately 15 degrees C for dimyristoyl- and 32 degrees C for dipalmitoylphosphatidylserine, and a pretransition is observed at approximately 6 degrees C (dimyristoylphosphatidylserine) and 21.5 degrees C (dipalmitoylphosphatidylserine), at J = 0.1. No titration of the transition is observed for the fully hydrated phosphate group down to pH less than or equal to 0.5, but it affinity for water binding decreases steeply at pH greater than or equal to 2.6. Increasing the NaCl concentration from 0.1 to 2.0 M increases the transition temperature of dimyristoyphosphatidylserine by approximately 8 degrees C at pH 7, by approximately 5 degrees at pH 13, and by approximately 0 degrees C at pH 1. These increases are attributed to the screening of the electrostatic titration-induced shifts in transition temperature. On a further increase of the NaCl concentration to 5.5 M, the transition temperature increases by an additional 9 degree C at pH 7, 13 degree C at pH 13, approximately 7 degree C in the fully hydrated state at pH 1, and approximately 4 and approximately 0 degree C in the two less hydrated states. These shifts are attributed to displacement of water of hydration by ion binding. From the salt dependence it is deduced that the transition temperature shift at the carboxyl titration can be accounted for completely by the surface charge and change in hydration of approximately 1 H2O/lipid, whereas that of the amino group titration arises mostly from other sources, probably hydrogen bonding. The shifts in pK (delta pK2 = 2.85, delta pK3 = 1.56) are consistent with a reduced polarity in the head-group region, corresponding to an effective dielectric constant epsilon approximately or equal to 30, together with surface potentials of psi congruent to -100 and -150 mV at the carboxyl and amino group pKs, respectively. The transition temperature of dimyristoylphosphatidylserine-water mixtures decreases by approximately 4 degree C each water/lipid molecule added, reaching a limiting value at a water content of approximately 9-10 H2O/lipid molecule.  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