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.
Intermolecular histone H4 interactions in core nucleosomes   总被引:1,自引:0,他引:1  
D G Chung  P N Lewis 《Biochemistry》1986,25(8):2048-2054
Chicken histone H4, labeled at methionine-84 with 1-N-pyrenyliodoacetamide, has been incorporated into a nucleosome-like particle with core length DNA and unmodified histones H2A, H2B, and H3. These synthetic nucleosomes exhibit properties very similar to those displayed by native particles and those labeled with other fluors. The emission spectrum of the pyrene-labeled nucleosome was characteristic of excited dimer (excimer) fluorescence, indicating that the single pyrene groups on the two H4 molecules are in close proximity in the reconstituted particle. Histone H4 was also labeled randomly at lysines with a group that contains two pyrene moieties separated by 12 A at most. Incorporation of this histone into nucleosome-like particles provides an excimer standard which does not depend on intermolecular interactions. The properties of the pyrene-containing nucleosome were examined as a function of ionic strength. It was found that the H4-H4 pyrene excimer fluorescence exhibited a cooperative disruption centered at 0.1 M NaCl which preceded increases in accessibility and environment polarity revealed by other fluors attached at the same site.  相似文献   

2.
Complexation between lysozyme and sodium poly(2-acrylamido-2-methylpropanesulfonate) (PAMPS) was studied by quasielastic light scattering, electrophoretic light scattering, fluorescence, and turbidimetry in electrolyte solution. These techniques show that complexation occurs at pH 9.6 in an ionic strength buffer of 0.25M NaCl + 25 mM Na2B4O7. At constant lysozyme concentration (Cpro). The structure of the complex depends on the polymer concentration. At low polymer concentration (relative to Cpro), an intrapolymer complex is formed. This intrapolymer complex aggregates to an interpolymer species upon increase in polymer concentration. Complex formation was also studied by fluorescence using pyrene-labeled PAMPS (Py-PAMPS). Energy transfer from singlet-excited tryptophan residues in lysozyme to the pyrene label occurs when the complex is formed. Fluorescence and turbidity data indicate that lysozyme interacts with Py-PAMPS preferentially at pyrene sites, which leads to static quenching of tryptophan fluorescence via energy transfer to the pyrene label. © 1995 John Wiley & Sons, Inc.  相似文献   

3.
A fluorescence-based detergent binding assay for protein hydrophobicity   总被引:2,自引:0,他引:2  
Protein hydrophobicity is often detected by binding of protein to micelles of a mild detergent. A new fluorescence method for detection of this binding is presented. The method is based on a long-range quenching of tryptophan fluorescence by energy transfer to a pyrene-labeled phospholipid probe incorporated into micelles of Brij 96. The method is rapid, simple, and requires only a few micrograms of protein. Strongest quenching is obtained when both pyrene probe and brominated Brij 96, a short-range quencher, are combined. To define the best assay conditions the physical properties and quenching behavior of micelles with or without these probes have been compared. It is shown that both quenchers accurately measure binding of model compounds and protein toxins to micelles. Comparison of quenching by the different probes can be used to derive information on tryptophan location relative to the micelle core.  相似文献   

4.
Properties of a specific glycolipid transfer protein from bovine brain   总被引:4,自引:0,他引:4  
A transfer protein specific for glycolipids has been isolated from bovine brain. As judged by sodium dodecyl sulfate (SDS) polyacrylamide gel electrophoresis, the protein is 68% pure and has a molecular weight of 20 000. Three different assays were employed to study the protein's specificity and glycolipid binding properties. The protein transferred several different neutral glycosphingolipids and ganglioside GM1 equally well, but failed to accelerate phosphatidylcholine or sphingomyelin intervesicular movement. The protein's ability to interact with glycolipids was strongly influenced by the physical properties of the matrix phospholipid in which the glycolipids reside. Both the phase state of the phospholipid matrix and bilayer curvature affected glycolipid intervesicular transfer rates. Protein binding to phospholipid vesicles containing either tritium-labeled or pyrene-labeled glucosylceramide could not be demonstrated by density gradient centrifugation or fluorescence energy transfer measurements, respectively. A specific association of the transfer protein for pyrene-labeled glucosylceramide was found when the fluorescence emission of the pyrene excimer-to-monomer ratio was measured suggesting that a portion of the fluorescent glycolipid was being sequestered from the phospholipid vesicles and was binding to the freely soluble protein.  相似文献   

5.
Here we have studied how the length of the pyrene-labeled acyl chain (n) of a phosphatidylcholine, sphingomyelin, or galactosylceramide affects the partitioning of these lipids between 1), gel and fluid domains coexisting in bovine brain sphingomyelin (BB-SM) or BB-SM/spin-labeled phosphatidylcholine (PC) bilayers or 2), between liquid-disordered and liquid-ordered domains in BB-SM/spin-labeled PC/cholesterol bilayers. The partitioning behavior was deduced either from modeling of pyrene excimer/monomer ratio versus temperature plots, or from quenching of the pyrene monomer fluorescence by spin-labeled PC. New methods were developed to model excimer formation and pyrene lipid quenching in segregated bilayers. The main result is that partition to either gel or liquid-ordered domains increased significantly with increasing length of the labeled acyl chain, probably because the pyrene moiety attached to a long chain perturbs these ordered domains less. Differences in partitioning were also observed between phosphatidylcholine, sphingomyelin, and galactosylceramide, thus indicating that the lipid backbone and headgroup-specific properties are not severely masked by the pyrene moiety. We conclude that pyrene-labeled lipids could be valuable tools when monitoring domain formation in model and biological membranes as well as when assessing the role of membrane domains in lipid trafficking and sorting.  相似文献   

6.
Structural models of F-actin suggest that three segments in actin, the DNase I binding loop (residues 38-52), the hydrophobic plug (residues 262-274) and the C-terminus, contribute to the formation of an intermolecular interface between three monomers in F-actin. To test these predictions and also to assess the dynamic properties of intermolecular contacts in F-actin, Cys-374 pyrene-labeled skeletal alpha-actin and pyrene-labeled yeast actin mutants, with Gln-41 or Ser-265 replaced with cysteine, were used in fluorescence experiments. Large differences in Cys-374 pyrene fluorescence among copolymers of subtilisin-cleaved (between Met-47 and Gly-48) and uncleaved alpha-actin showed both intra- and intermolecular interactions between the C-terminus and loop 38-52 in F-actin. Excimer band formation due to intermolecular stacking of pyrene probes attached to Cys-41 and Cys-265, and Cys-41 and Cys-374, in mutant yeast F-actin confirmed the proximity of these residues on the paired sites (to within 18 A) in accordance with the models of F-actin structure. The dynamic properties of the intermolecular interface in F-actin formed by loop 38-52, plug 262-274 and the C-terminus may account for the observed cross-linking of these sites with reagents < 18 A. The functional importance of actin filament dynamics was demonstrated by the inhibition of the in vitro motility in the Gln-41-Cys-374 cross-linked actin filaments.  相似文献   

7.
Apolipoprotein (apo) E is thought to undergo conformational changes in the N-terminal helix bundle domain upon lipid binding, modulating its receptor binding activity. In this study, site-specific fluorescence labeling of the N-terminal (S94) and C-terminal (W264 or S290) helices in apoE4 by pyrene maleimide or acrylodan was employed to probe the conformational organization and lipid binding behavior of the N- and C-terminal domains. Guanidine denaturation experiments monitored by acrylodan fluorescence demonstrated the less organized, more solvent-exposed structure of the C-terminal helices compared to the N-terminal helix bundle. Pyrene excimer fluorescence together with gel filtration chromatography indicated that there are extensive intermolecular helix-helix contacts through the C-terminal helices of apoE4. Comparison of increases in pyrene fluorescence upon binding of pyrene-labeled apoE4 to egg phosphatidylcholine small unilamellar vesicles suggests a two-step lipid-binding process; apoE4 initially binds to a lipid surface through the C-terminal helices followed by the slower conformational reorganization of the N-terminal helix bundle domain. Consistent with this, fluorescence resonance energy transfer measurements from Trp residues to acrylodan attached at position 94 demonstrated that upon binding to the lipid surface, opening of the N-terminal helix bundle occurs at the same rate as the increase in pyrene fluorescence of the N-terminal domain. Such a two-step mechanism of lipid binding of apoE4 is likely to apply to mostly phospholipid-covered lipoproteins such as VLDL. However, monitoring pyrene fluorescence upon binding to HDL(3) suggests that not only apoE-lipid interactions but also protein-protein interactions are important for apoE4 binding to HDL(3).  相似文献   

8.
The effect of alpha-tocopherol on the lipid fluidity of porcine intestinal brush-border membranes was studied using pyrene as a fluorescent probe. Addition of alpha-tocopherol to the medium decreased fluorescence intensity and lifetime, but increased the fluorescence polarization of pyrene-labeled membranes. beta-, gamma-, and delta-Tocopherols gave no appreciable effect on the fluorescence intensity and polarization of the complex. The apparent dissociation constant (3.1 +/- 0.12 microM) of the interaction of alpha-tocopherol with the membranes, estimated from the change in the fluorescence intensity with varying concentrations of alpha-tocopherol, was in good agreement with the concentration required to cause the half-maximal inhibition of lipid peroxidation of the membranes performed by incubation with 100 microM ascorbic acid and 10 microM Fe2+. Decrease of the slope in the thermal Perrin plot of the polarization of pyrene-labeled membranes by alpha-tocopherol suggests that the movement of pyrene molecules in the membranes is restricted by binding of the tocopherol. This interpretation was confirmed by an increased harmonic mean of the rotational relaxation time of the dye molecules in the membranes from 10.9 +/- 0.16 to 18.5 +/- 0.51 microseconds after addition of 25 microM alpha-tocopherol to the medium. The perturbation of lipid phase in the membranes induced by alpha-tocopherol was also suggested from a decreased quenching rate constant of pyrene fluorescence in the membranes for Tl+. Based on these results, the effect of alpha-tocopherol on the lipid fluidity of the membranes is discussed.  相似文献   

9.
Sahoo D  Narayanaswami V  Kay CM  Ryan RO 《Biochemistry》2000,39(22):6594-6601
Manduca sexta apolipophorin III (apoLp-III), an 18-kDa, monomeric, insect hemolymph apolipoprotein, is comprised of five amphipathic alpha-helices arranged as a globular bundle in the lipid-free state. Upon lipid binding, it is postulated that the bundle opens, exposing a continuous hydrophobic surface which becomes available for lipid interaction. To investigate lipid binding-induced helical rearrangements, we exploited the unique fluorescence characteristics of N-(1-pyrene)maleimide. Pyrene is a spatially sensitive extrinsic fluorescent probe, which forms excited-state dimers (excimers) upon close encounter with another pyrene molecule. Cysteine residues were introduced into apoLp-III (which otherwise lacks cysteine) at Asn 40 (helix 2) and/or Leu 90 (helix 3), creating two single-cysteine mutants (N40C-apoLp-III and L90C-apoLp-III) and N40C/L90C-apoLp-III, a double-cysteine mutant, which were labeled with pyrene maleimide. Pyrene-labeled N40C/L90C-apoLp-III, but not the pyrene-labeled single-cysteine mutants, exhibited strong excimer fluorescence in the lipid-free, monomeric state. Guanidine hydrochloride titration and temperature studies revealed a loss in excimer fluorescence, accompanied by a loss in the molar ellipticity of the protein. When apoLp-III interacts with phospholipid vesicles to form disklike complexes, a significant loss in excimer fluorescence was noted, indicating that the helices bearing the pyrene moieties diverge from each other. Pyrene excimer fluorescence was further employed to examine the relative orientation of lipid-bound apoLp-III molecules. Pyrene-labeled N40C- or L90C-apoLp-III displayed no excimer fluorescence in the disk complexes, while complexes prepared with an equal mixture of both single-labeled mutants did emit excimer fluorescence, indicating apoLp-III adopts a preferred nonrandom orientation around the perimeter of the bilayer disk. These studies establish pyrene excimer fluorescence as a useful spectroscopic tool to address intra- and intermolecular interactions of exchangeable apolipoproteins upon binding to lipid.  相似文献   

10.
The synthesis of a series of water-soluble galactopyranose-functionalized polystyrene-polyvinyl ether dendrigrafts and their characterization (in solution and thin solid deposits) have been achieved. The presence of external galactopyranose groups on dendritic polymers has been exploited to prepare dendrigraft-oligonucleotide conjugates using a simple one-step coupling procedure with amino-ended oligonucleotides (ODNs). Several parameters such as the peripherical density of hydrophilic branches, the polymerization degree of polystyrene or poly(hydroxyethyl vinyl ether) blocks, and the number of galactopyranose groups were tuned. A capture test with short labeled complementary ODNs (25 bases) confirmed the presence of covalently bound ODNs on various kinds of dendrigrafts. The ability of the dendritic polymers to enhance the sensitivity of enzyme-linked oligosorbent assay (ELOSA) diagnostic tests (detection of hepatitis B virus, DNA target of 2400 bases) was then evaluated, especially the influence of the macromolecular architecture and the impact of the structural parameters. The dendrigraft-ODN conjugate with the lower saccharide external density was found to lead to a very significant amplification of the fluorescence signal, corresponding to a limit of sensitivity of 10(9) DNA copies per milliliter (instead of 10(11) DNA copies per milliliter without using dendrigrafts). Conversely, the dendrigrafts exhibiting a very high number of branches and galactopyranose groups at their periphery were not able to induce a better sensitivity due to steric hindrance generated by the peripheral congestion on these polymers.  相似文献   

11.
Nanolipoprotein particles (NLPs) are nanometer‐scale discoidal particles that feature a phospholipid bilayer confined within an apolipoprotein “scaffold,” which are useful for solubilizing hydrophobic molecules such as drugs and membrane proteins. NLPs are synthesized either by mixing the purified apolipoprotein with phospholipids and other cofactors or by cell‐free protein synthesis followed by self‐assembly of the nanoparticles in the reaction mixture. Either method can be problematic regarding the production of homogeneous and monodispersed populations of NLPs, which also currently requires multiple synthesis and purification steps. Telodendrimers (TD) are branched polymers made up of a dendritic oligo‐lysine core that is conjugated to linear polyethylene glycol (PEG) on one end, and the lysine “branches” are terminated with cholic acid moieties that enable the formation of nanomicelles in aqueous solution. We report herein that the addition of TD during cell‐free synthesis of NLPs produces unique hybrid nanoparticles that have drastically reduced polydispersity as compared to NLPs made in the absence of TD. This finding was supported by dynamic light scattering, fluorescence correlation spectroscopy, and cryo transmission electron microscopy (Cryo‐EM). These techniques demonstrate the ability of TDs to modulate both the NLP size (6–30 nm) and polydispersity. The telodendrimer NLPs (TD‐NLPs) also showed 80% less aggregation as compared to NLPs alone. Furthermore, the versatility of these novel nanoparticles was shown through direct conjugation of small molecules such as fluorescent dyes directly to the TD as well as the insertion of a functional membrane protein.  相似文献   

12.
In this work we have applied a kinetic scheme derived from fluorescence kinetics of pyrene-labeled phosphatidylcholine in phosphatidylcholine membrane to explain the fluorescence quenching of 1-palmitoyl-2-(10-[pyrenl-yl]-sn-glycerol-3-phosphatidylchol ine (PPDPC) liposomes by tetracyanoquinodimethane (TCNQ). The scheme was also found to be applicable to neat PPDPC and the effect of the quencher could be attributed to certain steps of the proposed mechanism. The TCNQ molecules influence the fluorescence of pyrene moieties in PPDPC liposome in two ways. Firstly, an interaction between the quencher molecule and the pyrene monomer in the excited state quenches monomer fluorescence and effectively prevents the diffusional formation of the excimer. Secondly, an interaction between the quencher molecule and the excited dimer quenches the excimer fluorescence. The TCNQ molecule does not prevent the formation of the excimer in pyrene moieties aggregated in such a way that they require only a small rotational motion to attain excimer configuration. The diffusional quenching rate constant is calculated to be 1.0 x 10(8) M-1 s-1 for the pyrene monomer quenching and 1.3 x 10(7) M-1 s-1 for the pyrene excimer quenching. The diffusion constant of TCNQ is 1.5 x 10(-7) cm2 s-1 for the interaction radii of 0.8-0.9 nm. The TCNQ molecules are practically totally partitioned in the membrane phase.  相似文献   

13.
The effect of α-tocopherol on the lipid fluidity of porcine intestinal brush-border membranes was studied using pyrene as a fluorescent probe. Addition of α-tocopherol to the medium decreased fluorescence intensity and lifetime, but increased the fluorescence polarization of pyrene-labeled membranes. β-, γ-, and δ-Tocopherols gave no appreciable effect on the fluorescence intensity and polarization of the complex. The apparent dissociation constant (3.1 ± 0.12 μM) of the interaction of α-tocopherol with the membranes, estimated from the change in the fluorescence intensity with varying concentrations of α-tocopherol, was in good agreement with the concentration required to cause the half-maximal inhibition of lipid peroxidation of the membranes performed by incubation with 100 μM ascorbic acid and 10 μM Fe2+. Decrease of the slope in the thermal Perrin plot of the polarization of pyrene-labeled membranes by α-tocopherol suggests that the movement of pyrene molecules in the membranes is restricted by binding of the tocopherol. This interpretation was confirmed by an increased harmonic mean of the rotational relaxation time of the dye molecules in the membranes from 10.9 ± 0.16 to 18.5 ± 0.51 μs after addition of 25 μM α-tocopherol to the medium. The perturbation of lipid phase in the membranes induced by α-tocopherol was also suggested from a decreased quenching rate constant of pyrene fluorescence in the membranes for Tl+. Based on these results, the effect of α-tocopherol on the lipid fluidity of the membranes is discussed.  相似文献   

14.
The aldehydes present in acid-soluble type I collagen react with pyrenebutyrylhydrazine to form various types of complexes under different reaction conditions. These complexes exhibit one or more of three different pyrene fluorescence bands: monomer, excimer, and aggregate fluorescence. Collagen, whose aldehydes have been reduced with NaBH4, does not react with this fluorescent hydrazine, confirming that the hydrazine reacts specifically with aldehyde groups to form hydrazones. The absence of a reaction with pepsin-treated collagen also shows that the fluorescent labels are primarily in the nonhelical terminal telopeptides. Upon dialysis, the pyrene label bound to a saturated aldehyde in an α-chain is lost; whereas that bound to an unsaturated aldehyde remains on the protein. The pyrene monomer fluorescence in the β-chain of old collagen is stronger than that of young collagen. The formation of the pyrene excimer fluorescence implies the proximity of two pyrene molecules, probably attached to two adjacent aldehydes. Upon changing from acidic to neutral pH, both excimer and aggregate fluorescence bands disappear within a few seconds, revealing a very rapid alteration at the telopeptides.  相似文献   

15.
Deficiency of the vital muscle protein dystrophin triggers Duchenne/Becker muscular dystrophy, but the structure-function relationship of dystrophin is poorly understood. To date, molecular structures of three dystrophin domains have been determined, of which the N-terminal actin-binding domain (N-ABD or ABD1) is of particular interest. This domain is composed of two calponin-homology (CH) domains, which form an important class of ABDs in muscle proteins. A previously determined x-ray structure indicates that the dystrophin N-ABD is a domain-swapped dimer, with each monomer adopting an extended, open conformation in which the two CH domains do not interact. This structure is controversial because it contradicts functional studies and known structures of similar ABDs from other muscle proteins. Here, we investigated the solution conformation of the dystrophin N-ABD using a very simple and elegant technique of pyrene excimer fluorescence. Using the wild-type protein, which contains two cysteines, and the corresponding single-cysteine mutants, we show that the protein is a monomer in solution and is in a closed conformation in which the two CH domains seem to interact, as observed from the excimer fluorescence of pyrene-labeled wild-type protein. Excimer fluorescence was also observed in its actin-bound form, indicating that the dystrophin N-ABD binds to F-actin in a closed conformation. Comparison of the dystrophin N-ABD conformation with other ABDs indicates that the tandem CH domains in general may be monomeric in solution and predominantly occur in closed conformation, whereas their actin-bound conformations may differ.  相似文献   

16.
A method is reported for the synthesis of pyrene-labeled analogues of phosphatidylinositol 4-phosphate (Pyr-PIP) and phosphatidylinositol 4,5-biphosphate (Pyr-PIP2) from sn-2-(pyrenyl-decanoyl)phosphatidylinositol (Pyr-PI) using partially purified PI and PIP kinase preparations. Phosphorylation of Pyr-PI and Pyr-PIP was extensive (more than 50%) provided that the ATP concentration was high and that stabilizing agents such as sucrose and polyethylene glycol were present in the incubation medium. Pyr-PIP and Pyr-PIP2 were isolated by chromatography on immobilized neomycin. The identity of the products was established by thin-layer chromatography, UV-absorption spectroscopy, and spectrofluorometry. The pyrene excimer/monomer fluorescence technique revealed that, in contrast to Pyr-PI, Pyr-PIP and Pyr-PIP2 formed clusters in organic solvents. By use of the same technique for model membranes, it was shown that in phosphatidylcholine bilayers the collision frequency of the three fluorescent phosphoinositides decreased in the order PI greater than PIP greater than PIP2. Addition of Ca2+ at concentrations above 0.1 mM increased the collision frequency of Pyr-PIP2 and, to a much lesser extent, Pyr-PIP; Ca2+ had no effect on Pyr-PI.  相似文献   

17.
The effects of ionic strength on the conformation around the SH groups of the proteins and the lipid fluidity of porcine intestinal brush border membranes were studied using two fluorescent dyes, N-[7-dimethylamino-4-methylcoumarinyl]maleimide (DACM) and pyrene. The extent of DACM labeling to the SH groups of the membrane proteins was accelerated depending on the KCl concentrations in medium. A quenching study of DACM-labeled membranes with acrylamide showed that the proximity of the quencher to the fluorescence-labeled SH groups in the membrane proteins is increased with increasing ionic strength of medium. An implication of the conformational changes around SH groups in the membrane proteins with increase of ionic strength was also obtained from the stimulation of guanidine effect on the fluorescence parameters of DACM-labeled membranes by addition of KCl. On the other hand, the results of the quenching study with KI, excimer fluorescence, and polarization measurements of pyrene-labeled membranes suggested an increase of membrane fluidity on addition of KCl to medium. The temperature dependence of polarization of the complex strongly suggested that the rotational freedom of pyrene molecules embedded into the lipid layers of the membranes is increased by addition of KCl. In fact, the harmonic means of the rotational relaxation times of pyrene molecules in the membranes with and without 100 mM KCl were estimated to be about 2900 and 9000 ns at 25 degrees C, respectively. Based on these results, the salt-induced alterations of the conformation in the vicinity of the bound dyes of the membrane proteins and of the membrane fluidity are discussed.  相似文献   

18.
K Y Horiuchi  S Chacko 《Biochemistry》1988,27(22):8388-8393
Cysteine residues of caldesmon were labeled with the fluorescent reagent N-(1-pyrenyl)maleimide. The number of sulfhydryl (SH) groups in caldesmon was around 3.5 on the basis of reactivity to 5,5'-dithiobis(2-nitrobenzoate); 80% of the SH groups were labeled with pyrene. The fluorescence spectrum from pyrene-caldesmon showed the presence of excited monomer and dimer (excimer). As the ionic strength increased, excimer fluorescence decreased, disappearing at salt concentrations higher than around 50 mM. The labeling of caldesmon with pyrene did not affect its ability to inhibit actin activation of heavy meromyosin Mg-ATPase and the release of this inhibition in the presence of Ca2+-calmodulin. Tropomyosin induced a change in the fluorescence spectrum of pyrene-caldesmon, indicating a conformational change associated with the interaction between caldesmon and tropomyosin. The affinity of caldesmon to tropomyosin was dependent on ionic strength. The binding constant was 5 x 10(6) M-1 in low salt, and the affinity was 20-fold less at ionic strengths close to physiological conditions. In the presence of actin, the affinity of caldesmon to tropomyosin was increased 5-fold. The addition of tropomyosin also changed the fluorescence spectrum of pyrene-caldesmon bound to actin filaments. The change in the conformation of tropomyosin, caused by the interaction between caldesmon and tropomyosin, was studied with pyrene-labeled tropomyosin. Fluorescence change was evident when unlabeled caldesmon was added to pyrene-tropomyosin bound to actin. These data suggest that the interaction between caldesmon and tropomyosin on the actin filament is associated with conformational changes on these thin filament associated proteins. These conformational changes may modulate the ability of thin filament to interact with myosin heads.  相似文献   

19.
From the study of the kinetics of the fluorescence self-quenching of pyrene in multilamellar vesicles of 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) in the fluid phase we obtain the molecular diffusion parameters, diffusion coefficients, and activation energies for lateral diffusion for three probes: 1-palmitoyl-2-(1-pyrenedecanoyl)-sn-glycero-3-phosphocholine (py(10)-PC), pyrene, and 1-pyrenebutanoic acid (py-but). The experiments are done in a range of temperatures for which the reversibility of excimer formation is negligible and the probe/phospholipid ratios used are low, avoiding non-ideal mixing of the probe. The time-resolved fluorescence decays are, in all cases, accurately and consistently described by the two-dimensional (2D) formalism for bimolecular diffusion-controlled reactions. From the parameters obtained in this way we conclude that the primary step of the diffusion of the pyrene-labeled phospholipid involves the simultaneous jump of phospholipid and fluorophore moieties, and also that although in the case of py(10)-PC the pyrene from molecules pertaining to different layers may interact during the lifetime of the exited state, this is not the case for free pyrene.  相似文献   

20.
Recently, it was demonstrated that gamma-cyclodextrins (gamma-CDs) greatly accelerates transfer of hydrophobic pyrene-labeled and other fluorescent phospholipid derivatives from vesicles to cells in culture (). To understand better the characteristics of this process, we studied the interaction of gamma-CD with pyrene-labeled phosphatidylcholines (PyrPCs) using a variety of physical methods. Either one or both of the acyl chains of PC was labeled with a pyrene moiety (monoPyrPCs and diPyrPCs, respectively), and the length of the labeled chain(s) varied from 4 to 14 carbons. Fluorescent binding assays showed that the association constant decreases strongly with increasing acyl chain length. PyrPC/gamma-CD stoichiometry was 1:2 for the shorter chain species, but changed to 1:3 when the acyl chain length exceeded 8 (diPyrPCs) or 10 (monoPyrPCs) carbons. The activation energy for the formation of diPyr(10)PC/gamma-CD complex was high, i.e., +92 kJ/mol, indicating that the phospholipid molecule has to fully emerge from the bilayer before complex formation can take place. The free energy, enthalpy, and entropy of transfer of monoPyrPC from bilayer to gamma-CD complex were close to zero. The absorption, Fourier transform infrared, and fluorescence spectral measurements and lifetime analysis indicated that the pyrene moiety lies inside the CD cavity and is conformationally restricted, particularly when the labeled chain is short. The acyl chains of a PyrPC molecule seem to share a CD cavity rather than occupy different ones. The present data provide strong evidence that the ability of gamma-CD to enhance intermembrane transfer of pyrene-labeled phospholipids is based on the formation of stoichiometric complexes in the aqueous phase. This information should help in designing CD derivatives that are more efficient lipid carriers then those available at present.  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