Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 12 de 12
Filter
Add more filters










Publication year range
1.
Soft Matter ; 11(12): 2512-8, 2015 Mar 28.
Article in English | MEDLINE | ID: mdl-25682837

ABSTRACT

The structural dynamics of proteins is crucial to their biological functions. A precise and convenient method to determine the structural changes of a protein is still urgently needed. Herein, we employ fluorescence correlation spectroscopy (FCS) to track the structural transition of bovine serum albumin (BSA) in low concentrated cationic (cetyltrimethylammonium chloride, CTAC), anionic (sodium dodecyl sulfate, SDS), and nonionic (pentaethylene glycol monododecyl ether, C12E5 and octaethylene glycol monododecyl ether, C12E8) surfactant solutions. BSA is labelled with the fluorescence dye called ATTO-488 (ATTO-BSA) to obtain steady fluorescence signals for measurements. We find that the diffusion coefficient of BSA decreases abruptly with the surfactant concentration in ionic surfactant solutions at concentrations below the critical micelle concentration (CMC), while it is constant in nonionic surfactant solutions. According to the Stokes-Sutherland-Einstein equation, the hydrodynamic radius of BSA in ionic surfactant solutions amounts to ∼6.5 nm, which is 1.7 times larger than in pure water or in nonionic surfactant solutions (3.9 nm). The interaction between BSA and ionic surfactant monomers is believed to cause the structural transition of BSA. We confirm this proposal by observing a sudden shift of the fluorescence lifetime of ATTO-BSA, from 2.3 ns to ∼3.0 ns, in ionic surfactant solutions at the concentration below CMC. No change in the fluorescence lifetime is detected in nonionic surfactant solutions. Moreover, by using FCS we are also able to identify whether the structural change of protein results from its self-aggregation or unfolding.


Subject(s)
Serum Albumin, Bovine/chemistry , Surface-Active Agents/chemistry , Animals , Cattle , Fluorescence , Protein Conformation , Sodium Dodecyl Sulfate/chemistry , Spectrometry, Fluorescence
2.
Anal Chim Acta ; 855: 51-9, 2015 Jan 15.
Article in English | MEDLINE | ID: mdl-25542089

ABSTRACT

We determine the association constants for ligand-protein complex formation using the flow injection method. We carry out the measurements at high flow rates (F=1 mL min(-1)) of a carrier phase. Therefore, determination of the association constant takes only a few minutes. Injection of 1 nM of the ligand (10 µL of 1 µM concentration of the ligand solution) is sufficient for a single measurement. This method is tested and verified for a number of complexes of selected drugs (cefaclor, etodolac, sulindac) with albumin (BSA). We obtain K=4.45×10(3) M(-1) for cefaclor, K=1.00×10(5) M(-1) for etodolac and K=1.03×10(5) M(-1) for sulindac in agreement with the literature data. We also determine the association constants of 20 newly synthesized 3ß- and 3α-aminotropane derivatives with potential antipsychotic activity--ligands of 5-HT1A, 5-HT2A and D2 receptors with the albumin. Results of the studies reported here indicate that potential antipsychotic drugs bind weakly to the transporter protein (BSA) with K≈10(2)-10(3) M(-1). Our method allows measuring K in a wide range of values (10(2)-10(9) M(-1)). This range depends only on the solubility of the ligand and sensitivity of the detector.


Subject(s)
Pharmaceutical Preparations/metabolism , Serum Albumin, Bovine/metabolism , Animals , Antipsychotic Agents/metabolism , Cattle , Cefaclor/metabolism , Etodolac/metabolism , Ligands , Protein Binding , Sulindac/metabolism , Time Factors , Tropanes/metabolism
3.
Biosens Bioelectron ; 51: 8-15, 2014 Jan 15.
Article in English | MEDLINE | ID: mdl-23920090

ABSTRACT

Fluorescent double-stranded DNA (dsDNA) molecules labeled at both ends are commonly produced by annealing of complementary single-stranded DNA (ssDNA) molecules, labeled with fluorescent dyes at the same (3' or 5') end. Because the labeling efficiency of ssDNA is smaller than 100%, the resulting dsDNA have two, one or are without a dye. Existing methods are insufficient to measure the percentage of the doubly-labeled dsDNA component in the fluorescent DNA sample and it is even difficult to distinguish the doubly-labeled DNA component from the singly-labeled component. Accurate measurement of the percentage of such doubly labeled dsDNA component is a critical prerequisite for quantitative biochemical measurements, which has puzzled scientists for decades. We established a fluorescence correlation spectroscopy (FCS) system to measure the percentage of doubly labeled dsDNA (PDL) in the total fluorescent dsDNA pool. The method is based on comparative analysis of the given sample and a reference dsDNA sample prepared by adding certain amount of unlabeled ssDNA into the original ssDNA solution. From FCS autocorrelation functions, we obtain the number of fluorescent dsDNA molecules in the focal volume of the confocal microscope and PDL. We also calculate the labeling efficiency of ssDNA. The method requires minimal amount of material. The samples have the concentration of DNA in the nano-molar/L range and the volume of tens of microliters. We verify our method by using restriction enzyme Hind III to cleave the fluorescent dsDNA. The kinetics of the reaction depends strongly on PDL, a critical parameter for quantitative biochemical measurements.


Subject(s)
Biosensing Techniques/methods , DNA/analysis , Fluorescent Dyes/analysis , Spectrometry, Fluorescence/methods , Algorithms , Base Sequence , Biosensing Techniques/instrumentation , DNA Cleavage , Equipment Design , Kinetics , Molecular Sequence Data , Spectrometry, Fluorescence/instrumentation
4.
J Biol Chem ; 287(28): 23878-86, 2012 Jul 06.
Article in English | MEDLINE | ID: mdl-22573335

ABSTRACT

The self-assembly of the tubulin homologue FtsZ at the mid-cell is a critical step in bacterial cell division. We introduce dynamic light scattering (DLS) spectroscopy as a new method to study the polymerization kinetics of FtsZ in solution. Analysis of the DLS data indicates that the FtsZ polymers are remarkably monodisperse in length, independent of the concentrations of GTP, GDP, and FtsZ monomers. Measurements of the diffusion coefficient of the polymers demonstrate that their length is remarkably stable until the free GTP is consumed. We estimated the mean size of the FtsZ polymers within this interval of stable length to be between 9 and 18 monomers. The rates of FtsZ polymerization and depolymerization are likely influenced by the concentration of GDP, as the repeated addition of GTP to FtsZ increased the rate of polymerization and slowed down depolymerization. Increasing the FtsZ concentration did not change the size of FtsZ polymers; however, it increased the rate of the depolymerization reaction by depleting free GTP. Using transmission electron microscopy we observed that FtsZ forms linear polymers in solutions which rapidly convert to large bundles upon contact with surfaces at time scales as short as several seconds. Finally, the best studied small molecule that binds to FtsZ, PC190723, had no stabilizing effect on Caulobacter crescentus FtsZ filaments in vitro, which complements previous studies with Escherichia coli FtsZ and confirms that this class of small molecules binds Gram-negative FtsZ weakly.


Subject(s)
Bacterial Proteins/chemistry , Caulobacter crescentus/metabolism , Cytoskeletal Proteins/chemistry , Light , Scattering, Radiation , Algorithms , Bacterial Proteins/metabolism , Bacterial Proteins/ultrastructure , Cytoskeletal Proteins/metabolism , Cytoskeletal Proteins/ultrastructure , GTP Phosphohydrolases/chemistry , GTP Phosphohydrolases/metabolism , Guanosine Diphosphate/metabolism , Guanosine Diphosphate/pharmacology , Guanosine Triphosphate/metabolism , Guanosine Triphosphate/pharmacology , Kinetics , Microscopy, Electron, Transmission , Protein Multimerization/drug effects , Protein Stability/drug effects , Pyridines/pharmacology , Solutions/chemistry , Thiazoles/pharmacology
5.
Langmuir ; 26(5): 3562-8, 2010 Mar 02.
Article in English | MEDLINE | ID: mdl-19725563

ABSTRACT

Single-walled carbon nanotubes (SWNTs) were incorporated into a lyotropic liquid crystal (LLC) matrix formed by n-dodecyl octaoxyethene monoether (C(12)E(6)) at room temperature through spontaneous phase separation induced by nonionic hydrophilic polymer poly(ethylene glycol) (PEG). The quality of SWNTs/LLC composite was evaluated by polarized microscopy observations and small-angle X-ray scattering (SAXS) measurements. The results obtained clearly indicated that SWNTs have been successfully incorporated into the LLC matrix up to a considerable high content without destroying the LLC matrix, although interesting changes of the LLC matrix were also induced by SWNTs incorporation. By varying the ratio of PEG to C(12)E(6), the type of LLC matrix can be controlled from hexagonal phase to lamellar phase. Temperature was found to have a significant influence on the quality of SWNTs/LLC composite, and tube aggregation can be induced at higher temperature. When SWNTs were changed to multiwalled carbon nanotubes (MWNTs), they became difficult to be incorporated into LLC matrix because of an increase in the average tube diameter.


Subject(s)
Hydrophobic and Hydrophilic Interactions , Liquid Crystals/chemistry , Nanotubes, Carbon/chemistry , Polyethylene Glycols/chemistry , Microscopy, Polarization , Scattering, Small Angle , Surface-Active Agents/chemistry , Temperature , X-Ray Diffraction
6.
Langmuir ; 26(1): 34-40, 2010 Jan 05.
Article in English | MEDLINE | ID: mdl-19810715

ABSTRACT

Aggregate transitions in salt-free catanionic surfactant mixtures of tetradecyltrimethylammonium hydroxide (TTAOH)/fatty acid were investigated as a function of surfactant concentration and temperature. Lauric acid (LA), myristic acid (MA), and palmitic acid (PA) were chosen for the current study. The TTAOH/LA mixture exhibited rich phase behavior at room temperature. With increasing total surfactant concentration (c(T)), a bluish vesicular (L(alphav)) phase, an isotropic micellar (L(1)) phase, and a birefringent lamellar (L(alpha)) phase were observed. Between the L(alphav) phase and the L(1) phase, a narrow L(alpha)'/L(1) two-phase region was determined. With increasing temperature, a transition from the L(alpha) phase to the L(1) phase was induced at higher c(T) whereas at lower c(T) an opposite transition from the L(1) phase to the L(alphav) phase was noticed. Thus surprisingly, we observed bilayer-to-micelle and micelle-to-bilayer transitions in the same catanionic surfactant system, both induced by the temperature increase. Replacing LA by MA and PA caused a continuous increase in the average Krafft point of the mixture. The L(alphav)-phase region and phase-separated region become larger. Moreover, a single L(1)-phase region was absent within the investigated temperature range.

7.
Phys Chem Chem Phys ; 11(40): 9025-32, 2009 Oct 28.
Article in English | MEDLINE | ID: mdl-19812821

ABSTRACT

We measured the viscosity of poly(ethylene glycol) (PEG 6000, 12,000, 20,000) in water using capillary electrophoresis and fluorescence correlation spectroscopy with nanoscopic probes of different diameters (from 1.7 to 114 nm). For a probe of diameter smaller than the radius of gyration of PEG (e.g. rhodamine B or lyzozyme) the measured nanoviscosity was orders of magnitude smaller than the macroviscosity. For sizes equal to (or larger than) the polymer radius of gyration, macroscopic value of viscosity was measured. A mathematical relation for macro and nanoviscosity was found as a function of PEG radius of gyration, R(g), correlation length in semi-dilute solution, xi, and probe size, R. For R < R(g), the nanoviscosity (normalized by water viscosity) is given by exp(b(R/xi)a), and for R > R(g), both nano and macroviscosity follow the same curve, exp(b(R/xi)a), where a and b are two constants close to unity. This mathematical relation was shown to equally well describe rhodamine (of size 1.7 nm) in PEG 20,000 and the macroviscosity of PEG 8,000,000, whose radius of gyration exceeds 200 nm. Additionally, for the smallest probes (rhodamine B and lysozyme) we have verified, using capillary electrophoresis and fluorescence correlation spectroscopy, that the Stokes-Einstein (SE) relation holds, providing that we use a size-dependent viscosity in the formula. The SE relation is correct even in PEG solutions of very high viscosity (three orders of magnitude larger than that of water).


Subject(s)
Polyethylene Glycols/chemistry , Animals , Cattle , Electrophoresis, Capillary , Nanotechnology , Solutions , Spectrometry, Fluorescence , Viscosity
10.
J Phys Chem B ; 111(41): 11907-14, 2007 Oct 18.
Article in English | MEDLINE | ID: mdl-17887787

ABSTRACT

We studied mixing in the initially separated binary mixture of polystyrene/5CB liquid crystal and ternary mixtures of water/surfactant C12E5/polymer PEG system. In both systems the domains of one phase were characterized by a much higher viscosity than the solvent matrix. We demonstrated experimentally that during mixing these domains decrease their size linearly with time without a visible change of the optical contrast (i.e., without a rapid change of their compositions). Computer simulations and a theoretical model explain quantitatively our experimental observations.

11.
J Phys Chem B ; 110(20): 9869-75, 2006 May 25.
Article in English | MEDLINE | ID: mdl-16706441

ABSTRACT

A small-angle light scattering (SALS) technique together with optical microscopy observation are used to investigate phase separation kinetics in films of low molecular weight thermotropic liquid crystal (4-cyano-4'-n-octyl-biphenyl, 8CB) with flexible polymer (polystyrene, PS). The growth of domains is studied as a function of time, film thickness, and film composition. The light scattering results are correlated with the images obtained by optical microscopy observation. In this paper, we study the breaking of a bicontinuous network of polymer in liquid crystal into droplets and their further growth via the coalescence-induced coalescence mechanism. The appearance of droplets in the system leads to a strong scattering at small wave vectors, while the bicontinuous network gives a peak at a nonzero wave vector. Superposition of these scattering intensities leads to the appearance of a second peak in the full scattering intensity signal, when the bicontinuous network starts to break up into disjointed elongated domains. Finally, both peaks merge into a single peak, which moves quickly toward zero wave vectors, indicating a complete transformation of elongated domains into spherical droplets of variable size. We found that the separation process does not depend on the size of the system. Irrespective of the sample thickness, the network breaks into fragments always at the same time after temperature quench. On the basis of morphological analysis, we found that the average size of the droplets which formed from the network grows with time, t, as t(alpha), alpha = 0.9 +/- 0.1, in the isotropic phase and in the nematic phase.

12.
J Phys Chem B ; 109(34): 16252-62, 2005 Sep 01.
Article in English | MEDLINE | ID: mdl-16853066

ABSTRACT

We investigate the relaxation phenomena in a polymer (polystyrene)/liquid crystal (4-cyano-4'-n-octyl-biphenyl) system, in its homogeneous isotropic phase near the isotropic-isotropic, isotropic-nematic, and isotropic-smectic coexistence curve, using both polarized and depolarized photon correlation spectroscopy (PCS). We study this system for different polystyrene molecular weights (4750, 12 500, and 65 000 g/mol), different compositions (50, 40, 30, and 10% polystyrene (PS) by weight), and different temperatures close to phase boundaries. First of all, we determine the phase diagrams of this system for the different molecular weights. The shape of the phase diagrams strongly depends on the molecular weight. However, in all cases, at low temperatures, these systems separate into an almost pure liquid crystalline (LC) phase and polystyrene-rich phase. PCS measurements show that the relaxation processes in the homogeneous phase are not affected by the proximity of the nematic, or smectic, boundaries (even at a temperature of 0.1 degrees C above the phase separation in two phases). In polarized PCS experiments, we always see three relaxation processes well separated in time: one, very fast, with a relaxation time of the order of 10(-5) s; a second one with a relaxation time within the range 10(-2)-10(-3) s; and a last one, very slow, with a relaxation time of the order of 1 s. Both the fast and slow modes are independent of the wave vector magnitude, while the intermediate relaxation process is diffusive. In depolarized PCS experiments, the intermediate mode disappears and only the fast and slow relaxation processes remain, and they are independent of the magnitude of the wave vector. The diffusive mode is the classical diffusive mode, which is associated with the diffusion of polymer chains in all polymer solutions. The fast mode is due to the rotational diffusion of 4-cyano-4'-n-octyl-biphenyl (8CB) molecules close to polystyrene chains (transient network). Finally, we assign the slowest mode to reorientational processes of small aggregates of PS chains that are not dissolved in 8CB.

SELECTION OF CITATIONS
SEARCH DETAIL
...