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.
DNA reassociation kinetics using the phenol emulsion reassociation technique (PERT) [Kohne, D. E., Levison, S. A. & Byers, M. J. (1977) Biochemistry 16 , 5329–5341] has been investigated at high DNA concentrations using an endonuclease S1 assay of reaction progress. Apparent second-order rate constants fall on two intersecting straight lines when presented as a function of DNA concentrations on a log–log plot. In the low DNA concentration range, the rate constants drop about 10-fold when concentration increases 1000-fold. In the high DNA concentration range, the rate constants drop more than 10-fold when concentration increases 10-fold. The slopes of these lines are the same in different solvents and at different temperatures. The intersection between the lines occurs when the available catalytic surface is saturated. At high DNA concentrations, high-complexity heterologous denatured DNA apparently competes 2–4 times better for the surface than homologous DNA because it does not participate in a reassociation reaction. Native and partially native DNA molecules cannot compete with single-stranded DNA for a saturated surface. At high DNA concentrations, reactions using PERT become dependent on the single-strand DNA length. Increasing length lowers reassociation rates.  相似文献   

2.
We present an analysis of hybridization experiments on a DNA chip studied by surface plasmon resonance imaging. The reaction constants at various temperatures and for different probe lengths are obtained from Langmuir isotherms and hybridization kinetics. The melting curves from temperature scans are also obtained without any labeling of the targets. The effects of the probe length on the hybridization thermodynamics, deduced from the temperature dependence of the reaction constants as well as from the melting curves, suggest dispersion in the length of the hybridization segments of the probes accessible to the targets. Those are, however, sufficient to suggest efficient point mutation detection from temperature scans.  相似文献   

3.
Jamieson ER  Lippard SJ 《Biochemistry》2000,39(29):8426-8438
High-mobility group (HMG) domain proteins bind specifically to the major DNA adducts formed by the anticancer drug cisplatin and can modulate the biological response to this inorganic compound. Stopped-flow fluorescence studies were performed to investigate the kinetics of formation and dissociation of complexes between HMG-domain proteins and a series of 16-mer oligonucleotide probes containing both a 1,2-intrastrand d(GpG) cisplatin cross-link and a fluorescein-modified deoxyuridine residue. Rate constants, activation parameters, and dissociation constants were determined for complexes formed by HMG1 domain A and the platinated DNA probes. The sequence context of the cisplatin adduct modulates the value of the associative rate constant for HMG1 domain A by a factor of 2-4, contributing significantly to differences in binding affinity. The rates of association or dissociation of the protein-DNA complex were similar for a 71 bp platinated DNA analogue. Additional kinetic studies performed with HMG1 domain B, an F37A domain A mutant, and the full-length HMG1 protein highlight differences in the binding properties of the HMG domains. The stopped-flow studies demonstrate the utility of the fluorescein-dU probe in studying protein-DNA complexes. The kinetic data will assist in determining what role these proteins might play in the cisplatin mechanism of action.  相似文献   

4.
A mathematical model based on receptor-ligand interactions at a cell surface has been modified and further developed to represent heterogeneous DNA-DNA hybridization on a solid surface. The immobilized DNA molecules with known sequences are called probes, and the DNA molecules in solution with unknown sequences are called targets in this model. Capture of the perfectly complementary target is modeled as a combined reaction-diffusion limited irreversible reaction. In the model, there are two different mechanisms by which targets can hybridize with the complementary probes: direct hybridization from the solution and hybridization by molecules that adsorb nonspecifically and then surface diffuse to the probe. The results indicate that nonspecific adsorption of single-stranded DNA on the surface and subsequent two-dimensional diffusion can significantly enhance the overall reaction rate. Heterogeneous hybridization depends strongly on the rate constants for DNA adsorption/desorption in the non-probe-covered regions of the surface, the two-dimensional (2D) diffusion coefficient, and the size of probes and targets. The model shows that the overall kinetics of DNA hybridization to DNA on a solid support may be an extremely efficient process for physically realistic 2D diffusion coefficients, target concentrations, and surface probe densities. The implication for design and operation of a DNA hybridization surface is that there is an optimal surface probe density when 2D diffusion occurs; values above that optimum do not increase the capture rate. Our model predicts capture rates in agreement with those from recent experimental literature. The results of our analysis predict that several things can be done to improve heterogeneous hybridization: 1) the solution phase target molecules should be about 100 bases or less in size to speed solution-phase and surface diffusion; 2) conditions should be created such that reversible adsorption and two-dimensional diffusion occur in the surface regions between DNA probe molecules; 3) provided that 2) is satisfied, one can achieve results with a sparse probe coverage that are equal to or better than those obtained with a surface totally covered with DNA probes.  相似文献   

5.
A simple, direct method for the detection of DNA–protein interaction was developed with electrochemical methods. Single-stranded DNA (ss-DNA) probes were prepared through the chemical bonding of an oligonucleotide to a polymer film bearing carboxylic acid groups, and double-stranded DNA (ds-DNA) probes were prepared through hybridization of the complementary sequence DNA on the ss-DNA probe. Impedance spectroscopy and differential pulse voltammetry (DPV) distinguished the interaction between the DNA probes with mouse Purβ (mPurβ), an ss-DNA binding protein, and with Escherichia coli MutH, a ds-DNA binding protein. Impedance spectra obtained before and after the interaction of DNA probes with these proteins clearly showed the sequence-specific ss-DNA preference of mPurβ and the sequence-specific ds-DNA preference of MutH. The concentration dependence of proteins on the response of the DNA probes was also investigated, and the detection limits of MutH and mPurβ were 25 and 3 μg/ml, respectively. To confirm the impedance results, the variation of the current oxidation peak of adenine of the DNA probe was monitored with DPV. The formation constants of the complexes formed between the probe DNA and the proteins were estimated based on the DPV results.  相似文献   

6.
D. L. Weaver 《Biopolymers》1984,23(4):675-694
The diffusion–collision model of protein folding has been solved exactly for a three-microdomain protein subunit. Numerical analysis shows that the exact kinetics may be excellently approximated in all cases studied by a standard chemical kinetics approach with the forward rate constants calculated from the mean folding time formula found previously.  相似文献   

7.
The hybridization kinetics for a series of designed 25mer probe–target pairs having varying degrees of secondary structure have been measured by UV absorbance and surface plasmon resonance (SPR) spectroscopy in solution and on the surface, respectively. Kinetic rate constants derived from the resultant data decrease with increasing probe and target secondary structure similarly in both solution and surface environments. Specifically, addition of three intramolecular base pairs in the probe and target structure slow hybridization by a factor of two. For individual strands containing four or more intramolecular base pairs, hybridization cannot be described by a traditional two-state model in solution-phase nor on the surface. Surface hybridization rates are also 20- to 40-fold slower than solution-phase rates for identical sequences and conditions. These quantitative findings may have implications for the design of better biosensors, particularly those using probes with deliberate secondary structure.  相似文献   

8.
Fabrication of DNA microarrays using unmodified oligonucleotide probes   总被引:14,自引:0,他引:14  
Call DR  Chandler DP  Brockman F 《BioTechniques》2001,30(2):368-72, 374, 376 passim
Microarrays printed on glass slides are often constructed by covalently linking oligonucleotide probes to a derivatized surface. These procedures typically require relatively expensive amine- or thiol-modified oligonucleotide probes that add considerable expense to larger arrays. We describe a system by which unmodified oligonucleotide probes are bound to either nonderivatized or epoxy-silane-derivatized glass slides. Biotinylated PCR products are heat denatured, hybridized to the arrays, and detected using an enzymatic amplification system. Unmodified probes appear to detach from the slide surface at high pH (> 10.0), suggesting that hydrogen bonding plays a significant role in probe attachment. Regardless of surface preparation, high temperature (up to 65 degrees C) and low ionic strength (deionized water) do not disturb probe attachment; hence, the fabrication method described here is suitable for a wide range of hybridization stringencies and conditions. We illustrate kinetics of room temperature hybridizations for probes attached to nonderivatized slides, and we demonstrate that unmodified probes produce hybridization signals equal to amine-modified, covalently bound probes. Our method provides a cost-effective alternative to conventional attachment strategies that is particularly suitable for genotyping PCR products with nucleic acid microarrays.  相似文献   

9.
Wang Y  Wang X  Guo SW  Ghosh S 《BioTechniques》2002,32(6):1342-1346
We derived a theoretical model that explains certain biases observed in the two-color microarray hybridization experiments reported in the literature. We show that true competition is achieved only when the hybridization kinetics of the two differentially labeled probes are the same. If the hybridization kinetics of the two differentially labeled probes is different, which can occur when the labeling and hybridization conditions for the two probes are dissimilar, then differential expression observed becomes a function of the amount of the target (i.e., DNA spotted on the slide). We use this model to validate the microarray methodology by determining the differential expression of four select Arabidopsis genes and two human genes (beta-actin and GAPDH) as a function of the amount of target arrayed. We show through both modeling and experiments that the rate constants for Cy5- and Cy3-labeled probes are the same under our exrimental conditions. Therefore, the target concentrations need not greatly exceed the probe concentration. It is obvious from the data presented that a simple treatment of an individual hybridization rate calculation does notfully describe what is occuring in today's complex, multispecies experiments. The method of validation is easily implemented to ensure data reliability by two-color microarray.  相似文献   

10.
ESR spectra of membrane spin probes are conventionally used to obtain structural information. Here we show, for the first time, that when a membrane-soluble compound undergoes a chemical reaction, time-dependent changes in the ESR spectra of membrane spin probes can yield information about the kinetics of reaction. A benzoic acid ester, analog of the local anesthetic tetracaine, partitions between aqueous and membrane phases, causing changes in membrane structure as monitored by the ESR spectra of a probe. At alkaline pH, the lineshapes are time-dependent and the spectra go back to that in the absence of drug. The changes follow pseudo-first order kinetics. This effect is due to drug hydrolysis leading to water-soluble products, as confirmed by direct spectrophotometric measurements of the reaction. The pseudo-first order rate constants found by the latter method are in very good agreement with those calculated by ESR. The rate of hydrolysis decreases with increasing membrane concentration. This phenomenon accounts in part for the increased potency and toxicity of the more membrane-soluble local anesthetics.  相似文献   

11.
Dynamics of DNA condensation   总被引:6,自引:0,他引:6  
D Porschke 《Biochemistry》1984,23(21):4821-4828
The condensation of DNA induced by spermine and spermidine is investigated by equilibrium titrations and stopped-flow and field-jump experiments using scattered light detection. The spermine concentration required for the cooperative condensation process is measured at different DNA concentrations; these data are used to evaluate both the condensation threshold degree of spermine binding and the binding constant of spermine according to an excluded-site model. Stopped-flow measurements of the spermine-induced condensation demonstrate the existence of two processes: (1) A "fast" reaction is observed in the millisecond time range, when the reactant concentrations are around 1 microM; it is associated with a characteristic induction period and is assigned to the intramolecular condensation reaction. (2) A slow reaction with time constants of, e.g., 100 s strongly dependent upon both spermine and DNA concentrations is assigned to an intermolecular DNA association. The unusual time course of the intramolecular condensation reaction with the induction period provides evidence for a "threshold kinetics". During the induction period, spermine molecules are bound to DNA, but the degree of binding remains below the threshold value. As soon as the degree of ligand binding arrives at the threshold, the DNA is condensed in a relatively fast reaction. Model calculations of the spermine binding kinetics according to an excluded-site model demonstrate that the spermine molecules bound to DNA are mobile along the double helix. A comparison of the experimental data with the results of Monte Carlo simulations suggests a rate constant of approximately 200 s-1 for spermine movement by one nucleotide residue.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
We have investigated the kinetics of interaction of cationic fluorescent lipophiles (dyes) rhodamine 123, rhodamine 6G, tetramethyl rhodamine ethyl ester, safranine O, 1,1'-diethyloxacarbocyanine, 1,1'-diethyloxadicarbocyanine, and 1,1'-diethylthiadicarbocyanine iodide with isolated respiring rat-liver mitochondria (RLM). Dye flux across the RLM inner membrane was measured by following the kinetics of fluorescence signal change after mixing of dye and RLM. The time course of fluorescence was analysed in terms of a kinetic model of the binding and transport processes involved. The rate constants of dye influx and efflux were extracted from the observed effect on the apparent time constant of fluorescence change to equilibrium intensity upon mixing dye with increasing concentrations of RLM. From the influx rate constants obtained, the apparent permeability constants for dye influx (at zero potential) across the membrane were calculated and ranged from 3 to 140 x 10(-4) cm/s. The influx rate constant was found to be linearly related to relative dye lipophilicity, as predicted by the model. As another test of the model, from the ratio of the influx and efflux rate constants, the apparent trans-membrane potential, psi, was calculated and found generally to agree with reported values, but to depend on the lipophilicity of the dye used. Not predicted by the simple model was a dissymmtry observed in the influx and efflux time constants for fluorescence change to equilibrium intensity. Inferences are made relating to the utility of these dyes as probes of psi.  相似文献   

13.
Nucleic acid testing requires skilled personnel and expensive instrumentation. A method for the colorimetric detection of oligonucleotides that combines cellulose microparticles with biomolecular recognition is presented. DNA sequences from Trypanosoma brucei and dengue are used as model targets. Cellulose microparticles (≈20 µm) are bioactived by anchoring anti‐biotin antibodies via fusions that combine a carbohydrate‐binding module (CBM) with the ZZ fragment of protein A. Samples are prepared by incubating DNA probes immobilized on ≈14 nm gold nanoparticles (AuNPs) with biotin‐labeled targets and mixed with bioactive microparticles. The presence of unlabeled targets could also be probed by introducing a second, biotinylated DNA probe. The target:probe‐AuNP hybrids are mixed with and captured by the microparticles, which change color from white to red. Depletion of AuNPs from the liquid is also signaled by a decrease in absorbance at 525 nm. It was possible to detect targets with concentrations as low as 50 n m . In the presence of noncomplementary targets, microparticles remain white and the liquid remains red. The system is able to discriminate targets with a high degree of homology (≈53%). Overall, it is demonstrated that simple systems for the visual detection of nucleic acids can be set up by combining cellulose microparticles with biomolecular recognition agents based on CBMs and AuNPs.  相似文献   

14.
The aim of this study was to investigate the oxidation of two common fluorescent probes, dichlorodihydrofluorescein (DCFH2) and dihydrorhodamine (DHR), and their oxidized forms, dichlorofluorescein and rhodamine, by the radical products of peroxynitrite chemistry, *OH, NO2*, and CO3*-. At pH 8.0-8.2, rate constants for the interaction of carbonate radical with probes were estimated to be 2.6 x 10(8) x M(-1) s(-1) for DCFH2 and 6.7 x 10(8) M(-1) s(-1) for DHR. Nitrogen dioxide interacted more slowly than carbonate radical with these probes: the rate constant for the interaction between NO2* and DCFH2 was estimated as 1.3 x 10(7) M(-1) s(-1). Oxidation of DHR by nitrogen dioxide led to the production of rhodamine, but the kinetics of these reactions were complex. Hydroxyl radical interacted with both probes with rate constants close to the diffusion-controlled limit. We also found that oxidized forms of these fluorescent probes reacted rapidly with carbonate, nitrogen dioxide, and hydroxyl radicals. These data suggest that probe oxidation may often be in competition with reaction of the radicals with cellular antioxidants.  相似文献   

15.
Abstract

Two new 10-methoxydibenzo[b,h][1,6]naphthyridine-2-carboxamide derivatives (R1 and R2) have been synthesized and characterized using different spectral techniques. The binding of these probes with DNA was investigated using spectral (Electronic, fluorescence, 1H NMR and circular dichroism) and molecular docking studies. These probes exhibited a strong fluorescence around 440?nm upon excitation around 380?nm. Electronic and competitive fluorescence titration studies, in HEPES [(4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid)] buffer/dimethyl sulfoxide (pH 7.4) medium, suggest that these probes bind strongly to DNA, which is substantiated by 1H NMR study. The binding constants are calculated to be 5.3?×?107 and 6.8?×?106 M?1 for R1 and R2, respectively. From the results of spectral studies, it is proposed that the mechanism of binding of these probes with DNA is through minor groove binding mode, which is further confirmed by circular dichroism and molecular docking studies. Initial cell viability screening using MTT (3-[4,5-methylthiazol-2-yl]-2,5-diphenyl-tetrazolium bromide) assay shows that normal Vero cells are viable towards these probes at nano molar concentration, which is the concentration range employed in the present study for DNA staining (IC50 in the order of 0.023?mM). The enhancement in fluorescence intensity of these probes upon binding with DNA enables the staining of DNA in agarose gel in gel electrophoresis experiment. The sensitivity of these probes is comparable with that of ethidium bromide and DNA amounts as low as 4 nano gram are detectable.

Communicated by Ramaswamy H. Sarma  相似文献   

16.
Timothy M. Lohman 《Biopolymers》1983,22(7):1697-1713
We present a quantitative model for the irreversible dissociation kinetics of cooperatively bound nonspecific protein–nucleic acid complexes. The model assumes that the major pathway of dissociation is via singly contiguously bound protein that “peels” off the ends of clusters of bound protein. It should therefore be most applicable for proteins that bind nucleic acids with high cooperativity (w > 103). Furthermore, the model assumes that no redistribution of bound protein occurs during the time course of the dissociation. Solutions to the rate equations are presented for the entire time course of the dissociation. Under initial conditions such that the nucleic acid is less than fully saturated with protein, a single-exponential decay is predicted (if w is large). However, when the nucleic acid lattice is initially fully saturated, zero-order kinetics, corresponding to a constant rate of protein dissociation, is predicted. The experimental observation of zero-order dissociation kinetics in a cooperative protein–nucleic acid system is a good qualitative indicator for the dissociation mechanism discussed here. A discussion of the analysis of experimental data that enables one to extract molecular rate constants is presented. Furthermore, comparisons are made between the nonredistributing model presented here and Epstein's model [Epstein, I. R. (1979) Biopolymers 18 , 2037–2050] in which protein can translocate infinitely quickly while bound to the nucleic acid, and hence protein clusters redistribute during dissociation and maintain an equilibrium distribution on the nucleic acid at all times.  相似文献   

17.
Recently, several groups have developed green fluorescent protein (GFP)-based Ca(2+) probes. When applied in cells, however, these probes are difficult to use because of a low signal-to-noise ratio. Here we report the development of a high-affinity Ca(2+) probe composed of a single GFP (named G-CaMP). G-CaMP showed an apparent K(d) for Ca(2+) of 235 nM. Association kinetics of Ca(2+) binding were faster at higher Ca(2+) concentrations, with time constants decreasing from 230 ms at 0.2 microM Ca(2+) to 2.5 ms at 1 microM Ca(2+). Dissociation kinetics (tau approximately 200 ms) are independent of Ca(2+) concentrations. In HEK-293 cells and mouse myotubes expressing G-CaMP, large fluorescent changes were observed in response to application of drugs or electrical stimulations. G-CaMP will be a useful tool for visualizing intracellular Ca2+ in living cells. Mutational analysis, together with previous structural information, suggests the residues that may alter the fluorescence of GFP.  相似文献   

18.
A microtiter-based assay system is described in which DNA hairpin probes with dangling ends and single-stranded, linear DNA probes were immobilized and compared based on their ability to capture single-strand target DNA. Hairpin probes consisted of a 16 bp duplex stem, linked by a T2-biotin·dT-T2 loop. The third base was a biotinylated uracil (UB) necessary for coupling to avidin coated microtiter wells. The capture region of the hairpin was a 3′ dangling end composed of either 16 or 32 bases. Fundamental parameters of the system, such as probe density and avidin adsorption capacity of the plates were characterized. The target DNA consisted of 65 bases whose 3′ end was complementary to the dangling end of the hairpin or to the linear probe sequence. The assay system was employed to measure the time dependence and thermodynamic stability of target hybridization with hairpin and linear probes. Target molecules were labeled with either a 5′-FITC, or radiolabeled with [γ-33P]ATP and captured by either linear or hairpin probes affixed to the solid support. Over the range of target concentrations from 10 to 640 pmol hybridization rates increased with increasing target concentration, but varied for the different probes examined. Hairpin probes displayed higher rates of hybridization and larger equilibrium amounts of captured targets than linear probes. At 25 and 45°C, rates of hybridization were better than twice as great for the hairpin compared with the linear capture probes. Hairpin–target complexes were also more thermodynamically stable. Binding free energies were evaluated from the observed equilibrium constants for complex formation. Results showed the order of stability of the probes to be: hairpins with 32 base dangling ends > hairpin probes with l6 base dangling ends > 16 base linear probes > 32 base linear probes. The physical characteristics of hairpins could offer substantial advantages as nucleic acid capture moieties in solid support based hybridization systems.  相似文献   

19.
Reactive oxygen species (ROS) such as superoxide anion, hydroxyl ((*)OH), peroxyl, and alkoxyl radicals may attack biological macromolecules giving rise to oxidative stress-originated diseases. Since (*)OH is very short-lived, secondary products resulting from (*)OH attack to various probes are measured. Although the measurement of aromatic hydroxylation with HPLC/electrochemical detection is more specific than the low-yield TBARS test, it requires sophisticated instrumentation. As a more convenient and less costly alternative, we used p-aminobenzoate, 2,4- and 3,5-dimethoxybenzoate probes for detecting hydroxyl radicals generated from an equivalent mixture of Fe(II)+EDTA with hydrogen peroxide. The produced hydroxyl radicals attacked both the probe and the water-soluble antioxidants in 37 degrees C-incubated solutions for 2h. The CUPRAC (i.e., our original method for total antioxidant capacity assay) absorbance of the ethylacetate extract due to the reduction of Cu(II)-neocuproine reagent by the hydroxylated probe decreased in the presence of (*)OH scavengers, the difference being proportional to the scavenging ability of the tested compound. A rate constant for the reaction of the scavenger with hydroxyl radical can be deduced from the inhibition of color formation. The second-order rate constants of the scavengers were determined with competition kinetics by means of a linear plot of A(0)/A as a function of C(scavenger)/C(probe), where A(0) and A are the CUPRAC absorbances of the system in the absence and presence of scavenger, respectively, and C is the molar concentration of relevant species. The 2,4- and 3,5-dimethoxybenzoates were the best probes in terms of linearity and sensitivity. Iodide, metabisulfite, hexacyanoferrate(II), thiourea, formate, and dimethyl sulfoxide were shown by the modified CUPRAC assay to be more effective scavengers than mannitol, glucose, lysine, and simple alcohols, as in the TBARS assay. The developed method is less lengthy, more specific, and of a higher yield than the classical TBARS assay. The hydroxyl radical scavenging rate constants of ascorbic acid, formate, and hexacyanoferrate(II) that caused interference in other assays could be easily found with the proposed procedure.  相似文献   

20.
We have developed a new analytical method to detect multiple DNA simultaneously based on the biobarcoded CdSe/ZnS quantum dot (QD) and magnetic microparticle (MMP). It was demonstrated by using oligonucleotide sequences of 64 bases associated with human papillomavirus 16 and 18 L1 genes (HPV-16 and HPV-18) as model systems. This analytical system involves three types of probes, a MMP probe and two streptavidin-modified QD probes. The MMPs are functionalized with HPV-16 and HPV-18 captures DNA to form MMP probes. The QDs are conjugated with HPV-16 or HPV-18 probe DNA along with FAM- or Rox-labeled random DNA to form HPV-16 and HPV-18 QD probes, respectively. A one-step hybridization reaction was performed by mixing the MMP probes, HPV-16 and HPV-18 target DNA (T-16 and T-18), HPV-16 and HPV-18 QD probes. Afterwards, the hybrid-conjugated microparticles were separated by a magnet and heated to remove the MMPs. Finally, the detections of T-16 and T-18 were done by measuring fluorescence signals of FAM and Rox, respectively. Under the optimum conditions, the fluorescence intensity exhibited a good linear dependence on target DNA concentration in the range from 8 × 10?11 to 8 × 10?? M. The detection limit of T-16 is up to 7 × 10?11 M (3σ), and that of T-18 is 6 × 10?11 M. Compared with other biobarcode assay methods, the proposed method that QDs were used as the solid support has some advantages including shorter preparation time of QD probes, faster binding kinetics and shorter analytical time. Besides, it is simple and accurate.  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