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.
Amsacrine is a DNA intercalating agent which is active against a number of tumours in mice and is used for the treatment of leukaemia in humans. In its DNA-bound form, amsacrine efficiently quenches the fluorescence of ethidium. Fluorescence lifetime studies demonstrate two populations of DNA-bound ethidium. The first, whose fluorescence lifetime is constant at approx. 3 ns and whose proportion increases with increasing amsacrine binding ratio, may comprise molecules bound in close proximity to amsacrine. The second, whose fluorescence lifetime is longer and variable (10-24 ns) and whose proportion decreases with increasing amsacrine binding ratio, may comprise molecules three or more base-pairs away from ethidium. Studies with a number of derivatives of 9-anilinoacridine containing different anilino substituents suggest that the observed wide variation in quenching capacity is correlated with the magnitude of the substituent dipole moment in a particular direction. Consideration of the geometry of the DNA-binding complex indicates that the negative pole of a dipole established in the anilino ring is directed towards a positively charged site on the ethidium molecule. Quenching of ethidium fluorescence may therefore occur where an electron-transfer complex has formed between ethidium and amsacrine molecules. To ascertain whether electron-transfer complex formation is biologically important in the amsacrine series, ethidium quenching has been quantitated and compared with activity against a transplantable neoplasm in mice, the Lewis lung carcinoma. Compounds which strongly quench ethidium fluorescence are in general highly active antitumour agents. The results are discussed in terms of a model where amsacrine has both a DNA-binding and a protein-binding domain, the latter possibly interacting by formation of an electron-transfer complex. The most likely protein-binding domain is on the enzyme topoisomerase II, the target for its cytotoxic activity.  相似文献   

2.
The antitumor agent amsacrine, 4'-(9-acridinylamino)methanesulfon-m-anisidide (m-AMSA), when bound to double-stranded DNA, particularly poly(deoxyadenylicthymidylic acid), reduced the fluorescence of bound ethidium without physically displacing it from DNA. Fluorescence lifetime measurements showed that the reduction of fluorescence was not due to reduction of the lifetime of the excited state of ethidium. Rather, a proportion of the DNA-bound ethidium changed to a state where the fluorescence was highly quenched. Several other 9-anilinoacridine derivatives, and also 9-hydroxyellipticine, caused quenching of ethidium-DNA fluorescence, whereas 9-aminoacridine, proflavin, and ellipticine had no effect. Resonance energy transfer (F?rster transfer) is not responsible for the effect since there is no spectral overlap between the absorption spectrum of any of the agents and the fluorescence emission spectrum of ethidium. It is suggested that quenching may occur as a result of reversible formation of electron-transfer complexes between the intercalating drug and the excited state of ethidium.  相似文献   

3.
The mutagenicity of CI-921, the 4-methyl-5-(N-methyl)carboxamide derivative of the clinical antileukaemia agent, amsacrine, has been assessed using both bacterial and mammalian cells. CI-921 is distinguished from amsacrine in its high activity against some experimental tumours and is currently undergoing phase I clinical trial. Like 9-aminoacridine and amsacrine, CI-921 is mutagenic to the Salmonella typhimurium frameshift tester strain TA1537, but shows no sign of inducing base pair changes in strain TA100. In Chinese hamster cell culture, however, it differs from 9-aminoacridine in causing extensive chromosomal aberrations and an increase in mutations at the hypoxanthine-guanine phosphoribosyltransferase locus. It induces the formation of tightly packed and multilayered colonies in treated cultures of C3H/10T1/2 cells, but its action differs from that of benzo[a]pyrene, which induces type III fibroblastic multilayered colonies. Side-by-side comparison of the mutagenic properties of CI-921 and amsacrine showed no substantial differences at similar toxicity, suggesting that the increased lipophilicity and DNA-binding affinity of CI-921, which are thought to contribute to its increased antitumour activity, do not concomitantly increase the efficiency of in vitro mutagenesis or cell transformation.  相似文献   

4.
The spectroscopic properties and specificity of binding of a fluorescent quaternary amine, ethidium, with acetylcholine receptor-enriched membranes from Torpedo californica have been examined. Competition binding with [3H]phencyclidine in the presence of carbamylcholine showed that ethidium binds with high affinity to a noncompetitive inhibitor site (KD = 3.6 X 10(-7) M). However, in the presence of alpha-toxin, ethidium's affinity is substantially lower (KD approximately 1 X 10(-3) M). Ethidium was also found to enhance [3H]acetylcholine binding with a KD characteristic of ethidium binding to a high-affinity noncompetitive inhibitor site. These findings indicate that ethidium binds to an allosteric site which is regulated by agonist binding and can convert the agonist sites from low to high affinity. Fluorescence titrations of ethidium in the presence of carbamylcholine yielded a similar KD (2.5 X 10(-7) M) and showed an ethidium stoichiometry of one site/acetylcholine receptor monomer. Ethidium was completely displaced by noncompetitive inhibitors such as phencyclidine, histrionicotoxin, and dibucaine. The enhanced fluorescence lifetime of the bound species showed that the increased fluorescence intensity reflects a 13-fold increase in quantum yield for the complex compared to ethidium in buffer. Fractional dissociation of ethidium with phencyclidine produced a double-exponential fluorescence decay rate with lifetime components characteristic of ethidium free in solution and bound to the receptor. These data argue that the alterations in ethidium fluorescence elicited by other ligands is due to a change in the fraction of specifically bound ethidium rather than a change in quantum yield of a pre-existing ethidium-acetylcholine receptor complex. The extent of polarization indicates that bound ethidium is strongly immobilized. The magnitude of the quantum yield enhancement and the shifts of excitation and emission maxima of bound ethidium suggest that its binding site is within a hydrophobic domain with limited accessibility to the aqueous phase.  相似文献   

5.
We have previously reported that the binding of ethidium bromide to the nucleosome core particle results in a stepwise dissociation of the structure which involves the initial release of one copy each of H2A and H2B (McMurray & van Holde, 1986). In this report, we have examined the absorbance and fluorescence properties of intercalated and outside bound forms of ethidium bromide. From these properties, we have measured the extent of external, electrostatic binding of the dye versus internal, intercalation binding to the core particle, free from contribution by linker DNA. We have established that dissociation is induced by the intercalation mode of binding to DNA within the core particle DNA, and not by binding to the histones or by nonintercalative binding to DNA. The covalent binding of [3H]-8-azidoethidium to the core particle clearly shows that less than 1.0 adduct is formed per histone octamer over a wide range of input ratios. Simultaneously, analyses of steady-state fluorescence enhancement and fluorescence lifetime data from bound ethidium complexes demonstrate extensive intercalation binding. Combined analyses from steady-state fluorescence intensity with equilibrium dialysis or fluorescence lifetime data revealed that dissociation began when approximately 14 ethidium molecules are bound by intercalation to each core particle and less than 1.0 nonintercalated ion pair was formed per core particle.  相似文献   

6.
We report a highly specific fluorescence lifetime imaging microscopy (FLIM) method for monitoring epidermal growth factor receptor (EGFR) phosphorylation in cells based on fluorescence resonance energy transfer (FRET). EGFR phosphorylation was monitored using a green fluorescent protein (GFP)-tagged EGFR and Cy3-conjugated anti-phosphotyrosine antibodies. In this FRET-based imaging method, the information about phosphorylation is contained only in the (donor) GFP fluorescence lifetime and is independent of the antibody-derived (acceptor) fluorescence signal. A pixel-by-pixel reference lifetime of the donor GFP in the absence of FRET was acquired from the same cell after photobleaching of the acceptor. We show that this calibration, by acceptor photobleaching, works for the GFP-Cy3 donor-acceptor pair and allows the full quantitation of FRET efficiencies, and therefore the degree of exposed phosphotyrosines, at each pixel. The hallmark of EGFR stimulation is receptor dimerisation [1] [2] [3] [4] and concomitant activation of its intracellular tyrosine kinase domain [5] [6] [7]. Trans-autophosphorylation of the receptor [8] [9] on specific tyrosine residues couples the activated dimer to the intracellular signal transduction machinery as these phosphorylated residues serve as docking sites for adaptor and effector molecules containing Src homology 2 (SH2; reviewed in [10]) and phosphotyrosine-binding (PTB) [11] domains. The time-course and extent of EGFR phosphorylation are therefore important determinants of the underlying pathway and resulting cellular response. Our results strongly suggest that secondary proteins are recruited by activated receptors in endosomes, indicating that these are active compartments in signal transduction.  相似文献   

7.
We report the photophysical and electrochemical properties of phenol-pyrrolidino[60]fullerenes 1 and 2, in which the phenol hydroxyl group is ortho and para to the pyrrolidino group, respectively, as well as those of a phenyl-pyrrolidino[60]fullerene model compound, 3. For the ortho analog 1, the presence of an intramolecular hydrogen bond is supported by (1)H NMR and FTIR characterization. The redox potential of the phenoxyl radical-phenol couple in this architecture is 240 mV lower than that observed in the associated para compound 2. Further, the C(60) excited-state lifetime of the hydrogen-bonded compound 1 in benzonitrile is 260 ps, while the corresponding lifetime for 2 is identical to that of the model compound 3 at 1.34 ns. Addition of excess organic acid to a benzonitrile solution of 1 gives rise to a new species, 4, with an excited-state lifetime of 1.40 ns. In nonpolar aprotic solvents such as toluene, all three compounds have a C(60) excited-state lifetime of ~1 ns. These results suggest that the presence of an intramolecular H-bond in 1 poises the potential of phenoxyl radical-phenol redox couple at a value that it is thermodynamically capable of reducing the photoexcited fullerene. This is not the case for the para analog 2 nor is it the case for the protonated species 4. This work illustrates that in addition to being used as light activated electron acceptors, pyrrolidino fullerenes are also capable of acting as built-in proton-accepting units that influence the potential of an attached donor when organized in an appropriate molecular design.  相似文献   

8.
The mutagenicity of a series of derivatives of 9-anilinoacridine, including the clinical antitumour agent amsacrine, has been assessed using a bacterial frameshift tester strain (Salmonella typhimurium TA1537) and a yeast petite colony assay (Saccharomyces cerevisiae 5178B). The results have been compared with microbial mammalian cell cytotoxicity, DNA binding affinity and acridine base strength (pKa). Compounds containing strong electron donor substituents on the acridine ring, and which have a high acridine pKa, show minimal frameshift mutagenicity but are strong inducers of petite yeast mutants. Conversely, some compounds which have a high DNA binding constant but a significant proportion of uncharged form at neutral pH, show high frameshift mutagenicity but minimal induction of petite mutants. It is hypothesised that this inverse relationship arises from the presence of trans-membrane drug transport mechanisms which act to exclude some compounds, particularly strongly basic compounds from the cytoplasm and to concentrate them in mitochondria.  相似文献   

9.
Influence of isopropanol (iPrOH) on the structural dynamics of Thermomyces lanuginosa lipase (TLL) was studied by steady-state, time-resolved, and stopped-flow fluorescence spectroscopy, monitoring the intrinsic emission of Trp residues. The fluorescence of the four Trps of the wild-type enzyme report on the global changes of the whole lipase molecule. To monitor the conformational changes in the so-called "lid," an alpha-helical surface loop, the single Trp mutant W89m (W117F, W221H, W260H) was employed. Circular dichroism (CD) spectra revealed that iPrOH does not cause major alterations in the secondary structures of the wild-type TLL and W89m. With increasing [iPrOH], judged by the ratio of emission intensities at 350 nm and 330 nm, the average microenvironment of the Trps in the wild-type TLL became more hydrophobic, whereas Trp89 of W89m moved into a more hydrophilic microenvironment. Time-resolved fluorescence measurements revealed no major changes to be induced by iPrOH neither in the shorter fluorescence lifetime component (tau(1) = 0.5--1.2 ns) for the wild-type TLL nor in the longer fluorescence lifetime component (tau(2) = 4.8--6.0 ns) in the wild-type TLL and the W89m mutant. Instead, for W89m on increasing iPrOH from 25% to 50% the value for tau(1) increased significantly, from 0.43 to 1.5 ns. The shorter correlation time phi(1) of W89m had a minimum of 0.08 ns in 25% iPrOH. Judged from the residual anisotropy r(infinity) the amplitude of the local motion of Trp89 increased upon increasing [iPrOH] 10%. Stopped-flow fluorescence spectroscopy measurements suggested the lid to open within approximately 2 ms upon transfer of W89m into 25% iPrOH. Steady-state anisotropies and longer correlation times revealed increasing concentrations of iPrOH to result also in the formation of dimers as well as possibly also higher oligomers by TLL.  相似文献   

10.
G Nemecz  F Schroeder 《Biochemistry》1988,27(20):7740-7749
The fluorescent sterol delta 5,7,9(11),22-ergostatetraen-3 beta-ol (dehydroergosterol) was investigated as a cholesterol analogue to examine sterol domains in and spontaneous exchange of sterol between 1-palmitoyl-2-oleoylphosphatidylcholine (POPC) small unilamellar vesicles (SUV). Fluorescence lifetime, acrylamide quenching analyses, and intermembrane exchange kinetics were consistent with the presence of at least two sterol domains in POPC. Fluorescence lifetime was determined by phase and modulation fluorescence spectroscopy and analyzed by nonlinear least-squares as well as continuous distributional analyses. Both methods demonstrated that pure dehydroergosterol in POPC SUV had two lifetime components (C) and fractional intensities (F) near C1 = 0.851 ns (F1 0.96) and C2 = 2.668 ns (F2 0.004). In contrast to component C1, the center of lifetime distribution, fractional intensity, and peak width of dehydroergosterol lifetime component C2 was dependent on the polarity of the medium and vesicle curvature. The sterol domain corresponding to dehydroergosterol component C2 was preferentially quenched by acrylamide. Acrylamide quenching of dehydroergosterol fluorescence demonstrated that the two lifetime components of dehydroergosterol were not due to transbilayer sterol domains with different lifetimes. In a spontaneous exchange assay not requiring separation of donor and acceptor SUV, the lifetime component C2, but not C1, shifted to a shorter lifetime with altered distributional width. The kinetics of these lifetime and distributional width changes best fitted a two-exponential function, with a fast exchange rate constant K1 = 0.0325 min-1, t1/2 = 21.3 min, and a slow rate constant k2 = 0.00275 min-1, t1/2 = 261 min. The fast exchanging pool correlates with the longer lifetime component C2. These kinetics were confirmed both by dehydroergosterol exchange measured with fluorescence intensity and by [3H]cholesterol exchange. In summary, lifetime, distributional width, acrylamide quenching, and classical exchange assay data are consistent with the presence of at least two pools of sterol in POPC SUV.  相似文献   

11.
The presence of lipid domains in cellular membranes and their characteristic features are still an issue of dividing discussion. Several recent studies implicate lipid domains in plasma membranes of mammalian cells as short lived and in the submicron range. Measuring the fluorescence lifetime of appropriate lipid analogues is a proper approach to detect domains with such properties. Here, the sensitivity of the fluorescence lifetime of1-palmitoyl-2-[6-[(7-nitro-2-1,3-benzoxadiazol-4-yl)amino]-hexanoyl]-sn-glycero-3-phospholipid (C6-NBD-phospholipid) analogues has been employed to characterize lipid domains in giant unilamellar vesicles (GUVs) and the plasma membrane of mammalian cells by fluorescence lifetime imaging (FLIM). Fluorescence decay of C6-NBD-phosphatidylcholine is characterized by a short and long lifetime. For GUVs forming microscopically visible lipid domains the longer lifetime in the liquid disordered (ld) and the liquid ordered (lo) phase was clearly distinct, being approximately 7 ns and 11 ns, respectively. Lifetimes were not sensitive to variation of cholesterol concentration of domain-forming GUVs indicating that the lipid composition and physical properties of those lipid domains are well defined entities. Even the existence of submicroscopic domains can be detected by FLIM as demonstrated for GUVs of palmitoyloleoyl phosphatidylcholine/N-palmitoyl-d-sphingomyelin/cholesterol mixtures. A broad distribution of the long lifetime was found for C6-NBD-phosphatidylcholine inserted in the plasma membrane of HepG2 and HeLa cells centered around 11 ns. FLIM studies on lipid domains forming giant vesicles derived from the plasma membrane of HeLa cells may suggest that a variety of submicroscopic lipid domains exists in the plasma membrane of intact cells.  相似文献   

12.
The lifetimes of fluorescent components of matrix NADH in isolated porcine heart mitochondria were investigated using time-resolved fluorescence spectroscopy. Three distinct lifetimes of fluorescence were resolved: 0.4 (63%), 1.8 (30%), and 5.7 (7%) ns (% total NADH). The 0.4 ns lifetime and the emission wavelength of the short component were consistent with free NADH. In addition to their longer lifetimes, the remaining pools also had a blue-shifted emission spectrum consistent with immobilized NADH. On the basis of emission frequency and lifetime data, the immobilized pools contributed >80% of NADH fluorescence. The steady-state kinetics of NADH entering the immobilized pools was measured in intact mitochondria and in isolated mitochondrial membranes. The apparent binding constants (K(D)s) for NADH in intact mitochondria, 2.8 mM (1.9 ns pool) and >3 mM (5.7 ns pool), were on the order of the estimated matrix [NADH] (approximately 3.5 mM). The affinities and fluorescence lifetimes resulted in an essentially linear relationship between matrix [NADH] and NADH fluorescence intensity. Mitochondrial membranes had shorter emission lifetimes in the immobilized poo1s [1 ns (34%) and 4.1 ns (8%)] with much higher apparent K(D)s of 100 microM and 20 microM, respectively. The source of the stronger NADH binding affinity in membranes is unknown but could be related to high order structure or other cofactors that are diluted out in the membrane preparation. In both preparations, the rate of NADH oxidation was proportional to the amount of NADH in the long lifetime pools, suggesting that a significant fraction of the bound NADH might be associated with oxidative phosphorylation, potentially in complex 1.  相似文献   

13.
The noncovalent interaction of 2-aminodipyrido[1,2-a:3',2'-d]imidazole (Glu-P-2) and its derivatives, which are potent mutagens isolated from L-glutamic acid pyrolysate, with calf thymus DNA was studied by steady-state and nanosecond fluorescence spectroscopies. The fluorescence of these compounds exhibits static quenching by noncovalent interaction with DNA. Fluorescence lifetimes of the free and intercalated states of these compounds were determined to be 9-10 and 0.5-1 ns, respectively. The bisintercalative effect of the dimeric analogue of Glu-P-2, bis(Glu-P-2)spermine (2GP-SP), to DNA was also investigated. This 2GP-SP, which has two Glu-P-2 moieties at each end of spermine, indicates a strong intramolecular interaction exhibiting remarkable quenching of fluorescence spectrum and lifetime (tau = 3.5 ns) in the absence of DNA. In the presence of DNA, however, the 3.5-ns lifetime component of fluorescence disappeared, and a two-exponential decay of fluorescence (t = approximately 10 and 1.5 ns) was observed at a DNA concentration of more than approximately 0.001 mM P, while the solution containing a very dilute DNA concentration (less than or equal to 0.001 mM P) exhibits a three-component decay of fluorescence (1.5, 3.5, and approximately 10 ns). The potent bis intercalation of two moieties in 2GP-SP with an identical DNA molecule was suggested by the DNA-concentration dependence of these fluorescence lifetimes and their intensity.  相似文献   

14.
A study was conducted comparing the organization of chlorophyll during development of the photosynthetic apparatus in dark-grown and light-grown pine and barley. The rationale was that gymnosperms, but not angiosperms, have a capacity to synthesize chlorophyll in darkness. Seedlings of Pinus brutia were germinated and grown in darkness or under photoperiodic (day/night) conditions. The low-temperature (77 K) fluorescence spectra of newly-emerging dark-grown seedlings exhibited a single fluorescence band peaking at 678–679 nm, which decayed primarily with a ∼5.5 ns lifetime. Over the first few days of growth, the emission shifted to longer wavelengths and a subnanosecond lifetime component became prevalent. After several days of dark growth the emission spectrum and lifetime profile of the low temperature fluorescence came to resemble those of light-grown pine and barley. At room temperature, dark-grown pine showed little variable fluorescence, though addition of DCMU caused a substantial fluorescence rise. Illumination with moderate light for a few hours was sufficient to 'photoinduce' the appearance of normal variable fluorescence. At 77 K, DCMU-treated samples clearly showed a very long-lived (∼40 ns) fluorescence lifetime component in light-grown pine and barley. This component was undetectable in dark-grown pine. If, however, dark-grown samples were illuminated either before or after DCMU addition and then frozen to 77 K, the ∼40 ns lifetime component appeared at a fluorescence intensity similar to that in light-grown samples. These results are explained primarily in terms of photoactivation of the photosystem II (PSII) donor side. The temporary maintenance of PSII in an inactive, highly-quenched state is suggested to provide an available, yet protected precursor for active PSII.  相似文献   

15.
Pierre Sebban  Ismaël Moya 《BBA》1983,722(3):436-442
Fluorescence lifetime spectra of Rhodopseudomonas sphaeroides chromatophores have been measured at room temperature by phase fluorimetry at 82 MHz in order to investigate the heterogeneity of the emission. The total fluorescence was decomposed into two main components. A constant component, Fc, centered at 865 nm, represents about 50% of the total emission from dark-adapted chromatophores (Fo) and has a lifetime of 0.55 ns. A variable component is centered at 890 nm. Upon closing the reaction centers, 5-fold increases take place in both emission yield and lifetime of this component. In the dark-adapted state, its lifetime is about 50 ps and its contribution to the total fluorescence is 70% at 890 nm. In the presence of sodium dithionite, a long-lifetime component (τD ? 4 ns) is observed. This probably arises from radical pair recombination between P+ and I? (P, the primary electron donor, is a dimer of bacteriochlorophyll; I, the primary electron acceptor, is a molecule of bacteriopheophytin). Its spectrum is nearly identical to that of the variable component. This emission seems to be present also under nonreducing conditions, although with a much weaker intensity than when the electron acceptor quinone is prereduced.  相似文献   

16.
The fluorescence decay of 1-(4-trimethylammonium-phenyl)-6-phenyl-1,3,5-hexatriene (TMA-DPH) was used to study micro-heterogeneity of 1,2-dimyristoyl-3-sn-phosphatidylcholine (DMPC) liposomes and to characterize the effect of phosphatidic acid on the correlation between fluorescence microheterogeneity and membrane permeability. The fluorescence decay, measured using multifrequency phase fluorometry, has been analyzed either by using a model of discrete exponential components or a model of continuous distribution of lifetime values. Both analyses have shown that TMA-DPH decay is characterized by two components: a long one of about 9 ns and a short one of about 5 ns. In the gel phase, at variance with previous DPH studies, the short component was associated with a large fractional intensity. The distributional analysis showed changes of lifetime values and width in correspondence to the calorimetric transitions. The presence of egg phosphatidic acid increased both long lifetime values and distributional width. The use of TMA-DPH as a probe to evaluate membrane heterogeneity using the distributional width is discussed. The effect of phosphatidic acid on the membrane surface and in the hydrophobic core has been related to its structural properties and to its role in water penetration.  相似文献   

17.
G Weber  F J Farris 《Biochemistry》1979,18(14):3075-3078
Environmentally sensitive fluorescent probes involve two groups, an electron donor and an electron acceptor, attached to an aromatic ring system, and maximal effects may be expected when these groups are as far apart as feasible. The syntheisis, characterization, and spectroscopic properties of 6-propionyl-2-(dimethylamino)naphthalene (PRODAN), a compound that fulfills these conditions, are described. The maximum of emission is at 401 nm in cyclohexane solution and at 531 nm in water solution, indicating an increase of dipole moment of approximately 20 D units on excitation to the lowest singlet state. The effect of temperature upon the spectral distribution and the bandwidth of fluorescence of PRODAN in 1:1 complexes with albumin shows the existence of a dynamic relaxation process of the protein surroundings within th 2-4 ns of the fluorescence lifetime.  相似文献   

18.
1. Fluorescamine (4-phenylspiro[furan-2,(3)1'-phthalan]-3,3'-dione) reacts rapidly with Escherichia coli RNA polymerase and produces a fluorescent derivative which is inactivated to an extent dependent upon reagent concentration. Excess fluorescamine is rapidly hydrolysed. Reaction is with xi-amino gruops of lysine residues in all subunits as revealed by gel electrophoresis and fluorescence scanning. 2. The extent of inactivation and fluorescence yield are diminished in the presence of added template, a finding which provides evidence for the existence of reactive and essential amino groups which can be at least partially shielded by DNA in the binary complexes. The relative decrease of fluorescence is greatest in the betabeta' subunits. Holoenzyme and core enzyme show essentially the same behavior. 3. The inactivation of activity by fluorescamine is primarily at the level of initiation. Template binding and chain propagation are less affected. 4. The enzyme derivatized by fluorescamine shows an intense fluorescence with a peak at 490 nm and an excitation maximum at 390 nm. The fluorescence lifetime is in the range of 3-8 ns and the emission is highly polarized. In reactions carried out at high ionic strength the fluorescence yield is approximately double that at low ionic strength and insensitive to the presence of template. 5. Energy transfer is observed between the derivatized enzyme as donor and ethidium bromide as acceptor in the presence of template to which both the enzyme and intercalating dye are bound. The transfer efficiency is a function of the relative concentrations and of the conditions of reaction with fluorescamine. An average transfer distance of approx. 4-5 nm has been calculated suggesting a close proximity between bound polymerase and helical regions of the template.  相似文献   

19.
The fluorescence properties of the single tryptophanyl residue present in amicyanin from Thiobacillus versutus are very similar to those of azurin from Pseudomonas aeruginosa and other mononuclear blue copper proteins. The emission maximum is well structured and centered at 318 nm. The quantum yield is strongly affected by the presence of copper, the removal of which is accompanied by a more than sixfold increase in fluorescence, without change in shape. The fluorescence decay of holo-amicyanin is heterogeneous with a longer component of 5.7 ns and a shorter one of 0.7 ns accounting for 90% of the total emitting molecules. Copper-free amicyanin shows instead a single exponential decay (3.3 ns) of intrinsic fluorescence. This lifetime decreases as the temperature increases as does the longer lifetime component of holoamicyanin.  相似文献   

20.
The energy transfer and charge separation kinetics of a photosystem I (PS I) core particle of an antenna size of 100 chlorophyll/P700 has been studied by combined fluorescence and transient absorption kinetics with picosecond resolution. This is the first combined picosecond study of transient absorption and fluorescence carried out on a PS I particle and the results are consistent with each other. The data were analyzed by both global lifetime and global target analysis procedures. In fluorescence major lifetime components were found to be 12 and 36 ps. The shorter-lived one shows a negative amplitude at long wavelengths and is attributed to an energy transfer process between pigments in the main antenna Chl pool and a small long-wavelength Chl pool emitting around 720 nm whereas the longer-lived component is assigned to the overall charge separation lifetime. The lifetimes resolved in transient absorption are 7-8 ps, 33 ps, and [unk]1 ns. The shortest-lived one is assigned to energy transfer between the same pigment pools as observed also in fluorescence kinetics, the middle component of 33 ps to the overall charge separation, and the long-lived component to the lifetime of the oxidized primary donor P700+. The transient absorption data indicate an even faster, but kinetically unresolved energy transfer component in the main Chl pool with a lifetime <3 ps. Several kinetic models were tested on both the fluorescence and the picosecond absorption data by global target analysis procedures. A model where the long-wave pigments are spatially and kinetically connected with the reaction center P700 is favored over a model where P700 is connected more closely with the main Chl pool. Our data show that the charge separation kinetics in these PS I particles is essentially trap limited. The relevance of our data with respect to other time-resolved studies on PS I core particles is discussed, in particular with respect to the nature and function of the long-wave pigments. From the transient absorption data we do not see any evidence for the occurrence of a reduced Chl primary electron acceptor, but we also can not exclude that possibility, provided that reoxidation of that acceptor should occur within a time <40 ps.  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