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 125 毫秒
1.
2.
Daily turnover of cholesterol obtained by the balance method was compared to daily input rates calculated by input-output analysis in 43 experiments. The mean value of input rates for kinetic data of 10.1-16.4 weeks' duration (14 experiments) was 1.05 g/day vs. the chemical turnover of 0.94 g/day (difference 10.9 percent). For decay curves of 4.8-9.9 weeks' duration (29 experiments) the mean results were 1.67 g/day vs. 1.31 g/day, respectively (difference 20.1 percent). A combination of the balance method with input-output analysis is proposed to estimate the size of M (minimum value of the total exchangeable mass of cholesterol) in short-term experiments. Using this method, the analysis of curves of 10-12 weeks' duration showed a mean difference of 7.5 percent with the analysis of curves of 50-66 weeks' duration in 17 patients. However, because of considerable variations that can occur in individual cases, it is urged that a standard correction factor not be used, either in estimating turnover data or M from 10-12 weeks' kinetic data; rather, the proposed combined method will alert the investigator to the occurrence of discrepant results.  相似文献   

3.
The catabolism of intravenously injected 125I-labelled low density lipoproteins (LDL) was followed in normal miniature swine for 2 weeks. When compared with the two-exponential model, the decay curve of the plasma radioactivity associated with the LDL fraction was best described by a three-exponential model. In this system, the half-lives were 4.5 +/- 3.7, 19.7 +/- 6.6, and 127 +/- 70 h (mean of four studies). Assuming a kinetic model with metabolism of LDL in the rapidly equilibrating compartment and two slower equilibrating compartments (a model requiring three exponentials), the mean fractional catabolic rate for apo-LDL was calculated to be 0.015 h-1. Therefore, if at steady state, the synthetic rate for apo-LDL in the same pigs would be 5.6 +/- 4.1 mg/h. Different kinetic models using two or three exponentials would provide different values for the synthetic rate of apo-LDL. However, in view of the known existence of at least three major equilibrating pools for LDL in plasma, liver, and lymph, and in view of the present results, the kinetic model for LDL metabolism should be better represented by a three-exponential system.  相似文献   

4.
Total body turnover of cholesterol was studied in two patients with abetalipoproteinemia, a 32-year-old man and a 31-year-old woman. The patients received [14C]cholesterol intravenously, and the resulting specific activity-time curves (for 40 and 30 weeks, respectively) were fitted with a three-pool model. Parameters were compared with those from studies of cholesterol turnover in 82 normal and hyperlipidemic subjects. A three-pool model gave the best fit for the abetalipoproteinemic patients, as well as for the 82 previously studied subjects, suggesting general applicability of this model. Cholesterol production rates in the two abetalipoproteinemic subjects (0.82 and 0.89 g/day) were close to values predicted for persons of their body weight. Thus, total body turnover rate of cholesterol was quite normal in abetalipoproteinemia, confirming previous reports. Very low values (9.2 and 8.4 g) were found for M1, the size of the rapidly exchanging compartment pool 1, in the two abetalipoproteinemic subjects. These values were well below the values predicted (from the comparison study population) for normal persons of this size with low plasma cholesterol levels. For one patient, total body exchangeable cholesterol was very low, although not significantly below the predicted values for a person of his size. In the second patient, the observed estimate for total body exchangeable cholesterol was well within the range of values predicted for persons of her size with low to extremely low cholesterol levels.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
A fluorescence and radiolabel study of sterol exchange between membranes   总被引:2,自引:0,他引:2  
The fluorescent sterols delta 5,7,9(11),22-ergostatetraen-3 beta-ol (dehydroergosterol) and delta 5,7,9,(11)-cholestatrien-3 beta-ol (cholestatrienol) as well as [1,2-3H]cholesterol were utilized as cholesterol analogues to examine spontaneous exchange of sterol between 1-palmitoyl-2-oleoylphosphatidylcholine (POPC) small unilamellar vesicles (SUV). Exchange of fluorescent sterols was monitored at 24 degrees C by release from self-quenching of polarization from the time of mixing without separation of donor and acceptor vesicles. The polarization curve for 35 mol% sterol in POPC best fitted a two-exponential function, with a fast-exchange rate constant k1 = 0.0217 min-1, 1t1/2 = 32 min, size pool 1 = 12%, and a slow rate constant k2 = 2.91.10(-3) min-1, 2t1/2 = 238 min, size pool 2 = 88%. In addition to the above two exchangeable pools of sterol, the data were consistent with the presence of a slowly or nonexchangeable pool, 42% of total sterol, that was highly dependent on sterol content. These results were confirmed by simultaneous monitoring of [1,2-3H]cholesterol radioactivity and dehydroergosterol fluorescence intensity after separation of donor and acceptor vesicles by ion-exchange column chromatography. Thus, dehydroergosterol or cholestatrienol exchange as measured by fluorescence parameters (polarization and/or intensity) provides two new methods to follow cholesterol spontaneous exchange. These methods allow resolution and quantitation of a shorter exchange t1/2 near 30 min previously not reported. Thus, the cholesterol desorption rate from membranes may be faster than previously believed. In addition, the presence of a slowly non-exchangeable pool was confirmed.  相似文献   

6.
Dynamic light-scattering study on polymerization process of muscle actin   总被引:2,自引:0,他引:2  
Globular actin (G-actin) polymerizes into a fibrous form (F-actin) under physiological salt conditions. The polymerization process of muscle actin was studied by a dynamic light-scattering method. The intensity correlation functions G2(tau) of scattered light from a G-actin solution containing 2 mM Tris-HCl (pH 8.0) and 0.1 mM ATP were analyzed by a cumulant expansion method, and the translational diffusion coefficient was determined to be D = (8.07 +/- 0.10) X 10(-7) cm2/s at 20 degrees C. This D value gave a diameter of 5.3 nm for spherical G-actin including a hydration layer. Polymerization of 1-3 mg/ml G-actin in a solution containing 10 mM Tris-HCl (pH 8.0), 0.2 mM ATP and 60 mM KCl was followed by successive measurements of G2(tau) for a data accumulation period of 60-300 s/run. The time evolution of G2(tau) was analyzed by a least-squares fitting to the field correlation function of a multiexponential form g1(tau) = sigma iAi exp(-gamma i tau) with gamma 1 greater than gamma 2 greater than 3 greater than ..., and the static scattering intensity I(t) = mean value of I as a function of time t after initiation of polymerization was decomposed as I(t) = mean value of I sigma iAi. At the early stage of polymerization, a two-exponential fit gave results indicating that component 1 came from G-actin and component 2 from F-actin growing linearly with t. At the middle stage of polymerization, a three-exponential fit gave the results that component 1 came from G-actin and possibly its small oligomers, component 2 from polymers with a number-average length Ln of about 900 nm which was independent of t, and component 3 from 'ghosts' in dynamic light scattering in a semidilute regime. Component 3 was concluded to arise from restricted motions of polymers with lengths much longer than Ln in cages formed by polymers giving component 2, and a fragmentation-elongation process of F-actin was suggested to start at the middle stage of polymerization, resulting in the size redistribution of F-actin.  相似文献   

7.
We previously found a shorter surfactant disaturated phosphatidylcholine palmitate (DSPC-PA) half-life in infants with bronchopulmonary dysplasia (BPD) by using a single stable isotope tracer and simple formulas based on a one-exponential fit of the final portion of the enrichment decay curve. The aim of this study was to apply noncompartmental and compartmental analysis on the entire enrichment decay curve of DSPC-PA and to compare the kinetic data with our previous results. We analyzed 10 preterm newborns with BPD (gestational age 26 +/- 0.6 wk, weight 777 +/- 199 g) and 6 controls (gestational age 26 +/- 1.4 wk, weight 787 +/- 259 g). All took part in our previous study. Endotracheal 13C-labeled dipalmitoyl phosphatidylcholine was administered, and the 13C-enrichment of surfactant DSPC-PA was measured from serial tracheal aspirates by gas chromatography-mass spectrometry. Noncompartmental and compartmental models were numerically identified from the tracer-to-tracee ratio and kinetic parameters related to the accessible (pool accessible to sampling, likely to be the lung alveolar pool) and to the nonaccessible pools (pools not accessible to samplings, likely to be the intracellular storage pool) were estimated in the two study groups. Comparison was performed by Mann-Whitney test. A two-compartment model provided the most reliable assessment of DSPC-PA kinetics. In BPD vs. controls, mean +/- SE residence time of DSPC-PA in the accessible was 17.5 +/- 2.6 vs. 32.2 +/- 6.4 h (P < 0.05), whereas it was 49.7 +/- 3.5 vs. 54.4 +/- 3.9 h (NS, not significant) in the nonaccessible pool; DSPC-PA recycling was 0.26 +/- 0.05 vs. 0.43 +/- 0.04% (NS), respectively. A two-compartment model of surfactant DSPC-PA kinetics allowed a thorough assessment of DSPC-PA kinetics, including masses, synthesis, and fluxes between pools. The most important findings of this study are that in BPD infants DSPC-PA loss from the alveolar pool was higher and recycling through the intracellular pool lower than in controls.  相似文献   

8.
Stiffness behaviour of trabecular bone specimens   总被引:3,自引:0,他引:3  
Trabecular bone specimens were tested by non-destructive technique with the purpose of investigating stiffness behaviour and optimizing stiffness determination. Cylindrical specimens (n = 25) were loaded repetitively (0.1 Hz, 30 cycles) by axial compression to 50% of predicted ultimate strength and finally compressed to failure. Analyses of single compression curves showed increasing stiffness (E') until a stress level about 50% of ultimate stress followed by decreasing stiffness. Curve fit analysis of the elastic part of the compression curve showed the best fit, when a second order polynomial was used (r = 0.94, p less than 0.001). The stiffness determined non-destructively at the 25% level of ultimate strength increased significantly to the tenth loading cycle followed by a steady state. The precision of stiffness determination as an average of five consecutive measurements at steady state was E' +/- less than 5% (95% confidence limits). A reproducibility test by repetition of the test sequence after 3 h rest showed qualitatively the same stiffness behaviour. The variation of stiffness determination between the two test sequences was +/- 27% at the first loading cycle falling to +/- 12% at steady state.  相似文献   

9.
Short-term potentiation of ventilation (VSTP) may be observed in healthy subjects on sudden termination of an hypoxic stimulus. We hypothesized that the level of hypoxia preceding normoxia would modify the duration and magnitude of the ensuing ventilatory decay. Ten healthy adults were studied on two different occasions, during which they were randomly exposed to isocapnic 6 or 10% O2 for 60 s and then switched to an isocapnic normoxic gas mixture. Both hypoxic gases induced significant ventilatory responses, and mean peak minute ventilation before the isocapnic normoxic switch was higher in 6% O2 (P < 0.001). The fast time constant of the two-exponential equation representing the best fit for ventilatory decay was unaffected by the magnitude of the hypoxic stimulus. However, the slow time constant, which is considered to represent VSTP, was markedly prolonged in 6% compared with 10% O2 [106.7 +/- 11.3 vs. 38. 2 +/- 6.1 (SD) s, respectively; P < 0.0001]. This result indicates that VSTP is stimulus dependent. We conclude that the magnitude of hypoxia preceding a normoxic transient modifies VSTP characteristics. We speculate that the interdependence function of ventilatory stimulus and short-term potentiation is crucial for preservation of system stability during transitions from high to low ventilatory drives.  相似文献   

10.
Expressions for the partition function Q (T) of DNA hairpins are presented. Calculations of Q (T), in conjunction with our previously reported numerically exact algorithm [T. M. Paner, M. Amaratunga, M. J. Doktycz, and A. S. Benight (1990) Biopolymers, 29, 1715-1734], yield a numerical method to evaluate the temperature dependence of the transition enthalpy, entropy, and free energy of a DNA hairpin directly from its optical melting curve. No prior assumptions that the short hairpins melt in a two-state manner are required. This method is then applied in a systematic manner to investigate the stability of the six basepair duplex stem 5'-GGATAC-3' having four-base dangling single-strand ends with the sequences (XY)2, where X, Y = A, T, G, C, on the 5' end and a T4 loop on the 3' end. Results show that all dangling ends of the sample set stabilize the hairpin against melting. Increases in transition temperatures as great as 4.0 degrees C above the blunt-ended control hairpin were observed. The hierarchy of the hairpin transition temperatures is dictated by the identity of the first base of the dangling end adjoining the duplex in the order: purine greater than T greater than C. Calculated melting curves of every hairpin were fit to experimental curves by adjustment of a single parameter in the numerically exact theoretical algorithm. Exact fits were obtained in all cases. Experimental melting curves were also calculated assuming a two-state melting process. Equally accurate fits of all dangling-ended hairpin melting curves were obtained with the two-state model calculation. This was not the case for the melting curve of the blunt-ended hairpin, indicating the presence of a four-base dangling-end drives hairpin melting to a two-state process. Q (T) was calculated as a function of temperature for each hairpin using the theoretical parameters that provided calculated curves in exact agreement with the experimentally obtained optical melting curves. From Q (T), the temperature dependence of the transition enthalpy delta H, entropy delta S, and free energy delta G were calculated for every hairpin providing a quantitative assessment of the effects of dangling ends on hairpin thermodynamics. Comparisons of our results are made with those of the Breslauer group [M. Senior, R. A. Jones, and K. J. Breslauer (1988) Biochemistry 27, 3879-3885] on the T2 5' dangling-ended d(GC)3 duplexes.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

11.
The reduction in EPR signal intensity of nitroxide spin-labels by ascorbic acid has been measured as a function of time to investigate the immersion depth of the spin-labeled M2δ AChR peptide incorporated into a bicelle system utilizing EPR spectroscopy. The corresponding decay curves of n-DSA (n=5, 7, 12, and 16) EPR signals have been used to (1) calibrate the depth of the bicelle membrane and (2) establish a calibration curve for measuring the depth of spin-labeled transmembrane peptides. The kinetic EPR data of CLS, n-DSA (n=5, 7, 12, and 16), and M2δ AChR peptide spin-labeled at Glu-1 and Ala-12 revealed excellent exponential and linear fits. For a model M2δ AChR peptide, the depth of immersion was calculated to be 5.8? and 3? for Glu-1, and 21.7? and 19? for Ala-12 in the gel-phase (298K) and L(α)-phases (318K), respectively. The immersion depth values are consistent with the pitch of an α-helix and the structural model of M2δ AChR incorporated into the bicelle system is in a good agreement with previous studies. Therefore, this EPR time-resolved kinetic technique provides a new reliable method to determine the immersion depth of membrane-bound peptides, as well as, explore the structural characteristics of the M2δ AChR peptide.  相似文献   

12.
Chiang WH  Chen PY  Nien PC  Ho KC 《Steroids》2011,76(14):1535-1540
The indirect electrochemical oxidation method using bromine species in an organic media for measuring cholesterol was developed. The electrochemical behaviors of cholesterol were examined by cyclic voltammetry in a potential range of −0.5 to 2.0 V (vs. Ag/AgCl/saturated KCl). The polarization curve of the steady-state current in the applied potential range of 0-2.0 V is reported. The obtained kinetic parameters for the catalytic oxidation of cholesterol support the assumption that positive bromine species can be generated from bromine by undergoing two consecutive electrochemical oxidation steps. The positive bromine acts both as electron mediators and as electrocatalysts. Amperometric detection with an anodic current was investigated, and the calibration curve exhibited a linear relationship between the steady-state current and the concentration of cholesterol in the range of 30-200 μM, from which the sensitivity was calculated to be about 0.2 μA/cm2/μM. Moreover, the amperometric current followed Michaelis-Menten’s enzymatic model for cholesterol concentrations in the range of 30 μM to 5 mM, which can be applied for cholesterol rapid detection in processed foods.  相似文献   

13.
The effect of angiotensin-converting enzyme inhibition on function and structure of the aorta was studied in the Lewis polycystic kidney (LPK) rat model of cystic renal disease and Lewis controls. Pulse-wave velocity (PWV) was recorded under urethane anesthesia (1.3 g/kg ip) in mixed-sex animals aged 6 and 12 wk and in 12-wk-old animals treated with perindopril (3 mg·kg(-1)·day(-1) po) from age 6-12 wk. Tail-cuff systolic pressures were recorded over the treatment period. After PWV measurements, animals were euthanized and the aorta was removed for histomorphological and calcium analysis. Hypertension in LPK at 6 and 12 wk was associated with a shift of the PWV curve upward and to the right, indicating a decrease in aortic compliance, which was significantly reduced by perindopril. LPK demonstrated greater aortic calcification (6 wk: 123 ± 19 vs. 65 ± 7 and 12 wk: 406 ± 6 vs. 67 ± 6 μmol/g, P < 0.001, LPK vs. Lewis, respectively). This was reduced by treatment with perindopril (172 ± 48 μmol/g, 12 wk LPK P < 0.001). Medial cross-sectional area and elastic modulus/wall stress of the aorta were greater in LPK vs. Lewis control animals at 6 and 12 wk of age and showed an age-related increase that was prevented by treatment with perindopril (P < 0.001). Perindopril also ameliorated the degradation of elastin, increase in collagen content, and medial elastocalcinosis seen in 12-wk LPK. Overall, perindopril improved the structural and functional indices of aortic stiffness in the LPK rats, demonstrating a capacity for angiotensin-converting enzyme inhibition to limit vascular remodeling in chronic kidney disease.  相似文献   

14.
We used frequency-domain fluorometry to measure intensity and anisotropy decay of indole fluorescence in cyclohexane/ethanol mixtures at 20 degrees C. In 100% cyclohexane or 100% ethanol the intensity decay of indole appears to be a single exponential with decay times of 7.66 and 4.10 ns, respectively. In cyclohexane containing a small percentage of ethanol (up to 10%), we observed increased heterogeneity in intensity decay, resulting in a 10-fold increase in chi 2R for the single-exponential fit, as compared with the double-exponential model. We obtained comparable or better fits using unimodal Lorentzian and Gaussian lifetime distributions (two floating parameters) than for the two-exponential model (three floating parameters). We believe that the distribution of decay times reflects a range of indole solvation states in the dominately nonpolar solutions. This result suggests that a variety of hydrogen-bonding configurations could be one origin of the distributions of decay times observed for tryptophan emission from proteins. We also measured rotational diffusion of indole in cyclohexane, ethanol and its mixtures at 20 degrees C. The picosecond correlation times required that the mean decay times be decreased by acrylamide quenching (in ethanol) or energy transfer (in cyclohexane). In ethanol we observed nearly isotropic rotation of indole; in cyclohexane we obtained two correlation times of 17 and 73 ps. The shorter correlation time in cyclohexane appears to be due to the slip boundary condition, which was found to be progressively eliminated by small percentages of ethanol. Hence, hydrogen-bonding interactions appear to have a substantial effect on the rotational dynamics of indole.  相似文献   

15.
The curve for decay of 14C in rat liver protein labelled by injection of NaH14CO3 was analysed to obtain the average turnover rate of mixed liver protein. Three different methods of analysis were used. (1) Unlike decay curves from homogeneous proteins, the curve did not fit a single exponential, but a good fit was obtained with three exponentials. By assuming that the mixture contained three major components with different turnover rates, the calculated value for the average turnover rate (k) was close to 40% per day. (2) k was also calculated from the area under the decay curve, a method which makes no assumptions about the number of proteins in the mixture. This method also gave a value close to 40% per day. (3) It was shown empirically, both by simulation of decay of label in model mixtures of protein and with the decay curve measured in vivo, that k can be calculated from the time taken for the specific radioactivity to fall to 10% of its maximum value. This is an advantage, since the other two methods require the decay curve to be measured over a much longer period of time.  相似文献   

16.
Activated macrophages (M phi) appear responsible for at least part of the concomitant resistance in mice infected with Schistosoma mansoni. We found that as murine S. mansoni progressed from acute (8 to 12 wk of infection) to chronic (16 to 24 wk) stages, acquired resistance decreased (57% resistance to challenge with cercariae at 8 wk vs 28% by 24 wk, p less than 0.05), as did macrophage activation (21% +/- 2 killing of schistosomula by 8 wk M phi vs 8% +/- 2 for 24 wk M phi, p less than 0.01). T cells from the spleens of 8 wk-infected mice were capable of activating M phi from naive animals when stimulated with worm antigens (24% +/- 2 killing vs 8% +/- 2 induced by normal T cells, p less than 0.01); T cells obtained from 24 wk-infected mice did not activate M phi (13% +/- 2 killing). Furthermore, T cells from 24 wk-infected animals suppressed activation of M phi by 8 wk T cells. The addition of 10(5) 24 wk T cells to 3 X 10(5) antigen-stimulated 8 wk T cells reduced subsequent M phi killing from 27% +/- 4 to 13% +/- 2 (p less than 0.05). Week 24 T cells (3 X 10(5] reduced this additionally to 9% +/- 1 (p less than 0.01), a value no greater than that of unstimulated M phi. The subpopulation of T cells responsible for suppression of M phi activation was Lyt-2+1- nonadherent T cells. Cyclophosphamide treatment of chronically infected mice resulted in enhanced resistance (41%), M phi activation (18% +/- 1 killing), and T cell activation of naive M phi (10% +/- 1 killing). Thus, during chronic S. mansoni infection, resistance to reinfection wanes in parallel to and perhaps because of development of suppressor T cells that interfere with T-dependent M phi activation.  相似文献   

17.
Stopped-flow spectrometry and simple mixing techniques have been employed to investigate the detergent-induced dissociation of anthracycline antibiotics from natural and synthetic DNAs. Both daunomycin and nogalamycin dissociate more slowly from poly(dG-dC) than from poly(dA-dT) but the difference is much more marked for nogalamycin. With an equimolar mixture of poly(dG-dC) and poly(dA-dT), or with poly(dA-dC).poly(dG-dT), dissociation of nogalamycin occurs very slowly. In all cases the release of antibiotic from a synthetic polynucleotide is a one-step process following a single exponential. Dissociation of daunomycin, adriamycin and iremycin from calf thymus DNA is a more complex reaction which requires a two-exponential fit, in contrast to earlier reports, but differences between the behaviour of the three antibiotics are minor. Dissociation of nogalamycin from natural DNA requires a three-exponential fit, is in general far slower, and depends upon the base composition, the level of binding and the time allowed for the complex to equilibrate. It is concluded that sequence selectivity is minimal or lacking for daunomycin, whereas nogalamycin binding is sequence dependent and probably involves migration of the antibiotic between DNA binding sites. There is an inverse correlation between dissociation rate constants and antibacterial potency in simple tests.  相似文献   

18.
Stopped-flow spectrometry and simple mixing techniques have been employed to investigate the detergent-induced dissociation of anthracycline antibiotics from natural and synthetic DNAs. Both daunomycin and nogalamycin dissociate more slowly poly(dG-dC) than from poly(dA-dT), but the difference is much more marked for nogalamycin. With an equimolar mixture of poly(dG-dC) and poly(dA-dT), or with poly(dA-dC)·poly(dG-dT), dissociation of nogalamycin occurs very slowly. In all cases the release of antibiotic from a synthetic polynucleotide is a one-step process following a sinigle exponential. Dissociation of daunomycin, adrianmycin and iremycin from calf thymus DNA is a more complex reaction which requires a two-exponential fit, in contrast to earlier reports, but differences between the behaviour of the three antibotics are minor. Dissociation of nogalamycin from natural DNA requires a three-exponential fit, is in general far slower, and depends upon the base composition, the level of binding and the time allowed for the complex to equilibrate. It is concluded that sequence selectivity is minimal or lacking for daunomycin, whereas nogalamycin binding is sequence dependent and probably involves migration of the antibiotic between DNA binding sites. There is an inverse correlation between dissociation rate constants and antibacterial potency in simple tests.  相似文献   

19.
A N Zubov 《Tsitologiia》1980,22(10):1207-1213
Ionic currents through sodium channels of dialyzed mouse neuroblastoma N18 A-1 cells were measured under voltage clamp conditions. The PNa/PK ratio evaluated by reversal potential shifts was 10.4 +/- 0.7. Parameters of steady-state fast inactivation curves (h--V) and peak sodium conductance curves (gNa--V) were determined. The inactivation kinetics had usually a two-exponential time course. The internal perfusion of cells by trypsin and pronase caused a slowing-down of the sodium current falling phase, pronase being more specific in this respect. An external application of the purified scorpion toxin in concentration of 1.42 X 10(-7) M leads to a fast and sharp slowing-down of sodium inactivation. The same toxin in concentration of 5 X 10(-6) M, applied internally was quite unaffective. Experimental results demonstrate similarities in the main features between the sodium channels of neuroblastoma cells and those of other excitable cell membranes.  相似文献   

20.
The fluorescence kinetics of C-Phycocyanin in the monomeric, trimeric, and hexameric aggregation states has been measured as a function of the emission wavelength with picosecond resolution using the single-photon timing technique. All the decay curves measured at the various emission wavelengths were analyzed simultaneously by a global data analysis procedure. A sum of four exponentials was required to fit the data for the monomers and trimers. Only in the case of the hexamers, a three-exponential model function proved to be nearly sufficient to describe the experimental decays. The lifetime of those fluorescence components reflecting energy transfer decreased with increasing aggregation. This is due to the increased number of efficient acceptor molecules next to a donor in the higher aggregates. In all aggregates the shortest-lived component, ranging from 50 ps for monomer to 10 ps for hexamers, is observed as a decay term (positive amplitude) at short emission wavelength. At long emission wavelength it turns into a rise term (negative amplitude). The lifetime of a second ps-component ranges from 200 ps for monomers to 50 ps for hexamers. The long-lived (ns) fluorescence is inhomogeneous in monomers and trimers, showing two lifetimes of ~0.6 and 1.3 ns. The latter one carries the larger amplitude. The amplitudes of the kinetic components in the fluorescence decays are presented as time-resolved component spectra. A theoretical model has been derived to rationalize the observed fluorescence kinetics. Using symmetry arguments, it is shown that the fluorescence kinetics of C-Phycocyanin is expected to be characterized by three exponential kinetic components, independent of the aggregation state. An analytical expression is derived, which allows us to gain a detailed understanding of the origin of the different kinetic components and their associated time-resolved spectra. Numerical calculations of time-resolved spectra are compared with the experimental data.  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