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.
Scanning force spectroscopy was used to measure the mechanical properties of double stranded RNA molecules in comparison with DNA. We find that, similar to the B–S transition in DNA, RNA molecules are stretched from the assumed A′ conformation to a stretched conformation by applying a defined force (plateau force). The force depends on the G + C content of the RNA and is distinct from that required for the B–S transition of a homologous DNA molecule. After the conformational change, DNA can be further extended by a factor of 0.7 ± 0.2 (S-factor) before melting occurs and the binding of the molecule to the cantilever is finally disrupted. For RNA, the S-factor was higher (1.0 ± 0.2) and more variable. Experiments to measure secondary structures in single stranded RNA yielded a large number of different force-distance curves, suggesting disruption and stretching of various secondary structures. Oriented attachment of the molecules to the substrate, a defined pick-up point and an increased resolution of the instrument could provide the means to analyse RNA secondary structures by scanning force spectroscopy.  相似文献   

2.
Rapid construction of high-resolution physical maps requires accurate information about overlap between DNA clones and the size of gaps between clones or clone contigs. We recently developed a procedure termed ‘quantitative DNA fiber mapping’ (QDFM) to help construct physical maps by measuring the overlap between clones or the physical distance between non-overlapping contigs. QDFM is based on hybridization of non-isotopically labeled probes onto DNA molecules that were bound to a solid support and stretched homogeneously to ~2.3 kb/µm. In this paper, we describe the design of probes that bind specifically to the cloning vector of DNA recombinants to facilitate physical mapping. Probes described here delineate the most frequently used cloning vectors such as BACs, P1s, PACs and YACs. As demonstrated in representative hybridizations, vector-specific probes provide valuable information about molecule integrity, insert size and orientation as well as localization of hybridization domains relative to specifically-marked vector sequences.  相似文献   

3.
Here we describe bacterial genotyping by direct linear analysis (DLA) single-molecule mapping. DLA involves preparation of restriction digest of genomic DNA labeled with a sequence-specific fluorescent probe and stained nonspecifically with intercalator. These restriction fragments are stretched one by one in a microfluidic device, and the distribution of probes on the fragments is determined by single-molecule measurement of probe fluorescence. Fluorescence of the DNA-bound intercalator provides information on the molecule length. Because the probes recognize short sequences, they encounter multiple cognate sites on 100- to 300-kb-long DNA fragments. The DLA maps are based on underlying DNA sequences of microorganisms; therefore, the maps are unique for each fragment. This allows fragments of similar lengths that cannot be resolved by standard DNA sizing techniques to be readily distinguished. DNA preparation, data collection, and analysis can be carried out in as little as 5 h when working with monocultures. We demonstrate the ability to discriminate between two pathogenic Escherichia coli strains, O157:H7 Sakai and uropathogenic 536, and we use DLA mapping to identify microorganisms in mixtures. We also introduce a second color probe to double the information used to distinguish molecules and increase the length range of mapped fragments.  相似文献   

4.
When individual dsDNA molecules are stretched beyond their B-form contour length, they reveal a structural transition in which the molecule extends 1.7 times its contour length. The nature of this transition is still a subject of debate. In the first model, the DNA helix unwinds and combined with the tilting of the base pairs (which remain intact), results in a stretched form of DNA (also known as S-DNA). In the second model the base pairs break resulting effectively in two single-strands, which is referred to as force-induced melting. Here a combination of optical tweezers force spectroscopy with fluorescence microscopy was used to study the structure of dsDNA in the overstretching regime. When dsDNA was stretched in the presence of 10 nM YOYO-1 an initial increase in total fluorescence intensity of the dye–DNA complex was observed and at an extension where the dsDNA started to overstretch the fluorescence intensity leveled off and ultimately decreased when stretched further into the overstretching region. Simultaneous force spectroscopy and fluorescence polarization microscopy revealed that the orientation of dye molecules did not change significantly in the overstretching region (78.0°± 3.2°). These results presented here clearly suggest that, the structure of overstretched dsDNA can be explained accurately by force induced melting.  相似文献   

5.
DNA sequence analysis by oligonucleotide binding is often affected by interference with the secondary structure of the target DNA. Here we describe an approach that improves DNA secondary structure prediction by combining enzymatic probing of DNA by structure-specific 5′-nucleases with an energy minimization algorithm that utilizes the 5′-nuclease cleavage sites as constraints. The method can identify structural differences between two DNA molecules caused by minor sequence variations such as a single nucleotide mutation. It also demonstrates the existence of long-range interactions between DNA regions separated by >300 nt and the formation of multiple alternative structures by a 244 nt DNA molecule. The differences in the secondary structure of DNA molecules revealed by 5′-nuclease probing were used to design structure-specific probes for mutation discrimination that target the regions of structural, rather than sequence, differences. We also demonstrate the performance of structure-specific ‘bridge’ probes complementary to non-contiguous regions of the target molecule. The structure-specific probes do not require the high stringency binding conditions necessary for methods based on mismatch formation and permit mutation detection at temperatures from 4 to 37°C. Structure-specific sequence analysis is applied for mutation detection in the Mycobacterium tuberculosis katG gene and for genotyping of the hepatitis C virus.  相似文献   

6.
Kelly TJ 《Journal of virology》1975,15(5):1267-1272
The DNA molecule of the adenovirus 7-simian virus (SV) 40 hybrid, E46+, contains a single substitution of SV40 DNA located about 0.05 adenovirus 7 lengths from one end (arbitrarily designated the left end). The left to right direction in the SV40 DNA segment of the hybrid corresponds to the clockwise direction in the SV40 physical map. The left end point of the segment maps at about SV40 map position 0.50 ± 0.01 and the right end point at about SV40 map position 0.66 ± 0.01. The region between SV40 map positions 0.71 and 0.11 (clockwise) is deleted. Thus, the region between SV40 map position 0.50 and 0.66 (clockwise) is repeated in tandem.  相似文献   

7.
The effect of progressive denaturation of open circular molecules (component II) and supercoiled covalently closed circular molecules (component I) of rat liver mitochondrial DNA has been followed by heating in the presence of formaldehyde and examination in the electron microscope. After heating at 49°C, two, three, or four regions of strand separation were visible in 25% of the component II molecules. Comparisons of the patterns of distribution of these regions in individual molecules indicated that they occurred at at least three specific positions around the molecule. Also, these regions, which were assumed to be rich in adenine and thymine, were within a segment which was less than 50% of the length of the molecule. After heating at 50°C, up to 14 regions of strand separation were observed, but when comparisons were made no clear groupings were found. At 51°C, component II molecules were completely separated into a single-stranded circle and a single-stranded linear piece of similar length. Strand separation was accompanied by shortening of the molecule. At 70°C, single-stranded circles had a mean length of 2.7 µ, compared with 5.0 µ for native molecules. Progressive heating of component I molecules resulted first in conversion to an open circle (I') and then to a second supercoiled form (I'). Visualization of further denaturation products of component I was prevented by crosslinking of the molecule by formaldehyde at high temperatures.  相似文献   

8.
Stretching or aligning DNA molecules onto a surface by means of molecular combing techniques is one of the critical steps in single DNA molecule analysis. However, many of the current studies have focused on λ-DNA, or other large DNA molecules. There are very few studies on stretching methodologies for DNA molecules generated via PCR (typically smaller than 20 kb). Here we describe a simple method of stretching DNA molecules up to 18 kb in size on a modified glass surface. The very low background fluorescence allows efficient detection of single fluorescent dye labels incorporated into the stretched DNA molecules.  相似文献   

9.
SYBR Gold is a commonly used and particularly bright fluorescent DNA stain, however, its chemical structure is unknown and its binding mode to DNA remains controversial. Here, we solve the structure of SYBR Gold by NMR and mass spectrometry to be [2-(4-{[diethyl(methyl)ammonio]methyl}phenyl)-6-methoxy-1-methyl-4-{[(2Z)-3-methyl-1,3-benzoxazol-2-ylidene]methyl}quinolin-1-ium] and determine its extinction coefficient. We quantitate SYBR Gold binding to DNA using two complementary approaches. First, we use single-molecule magnetic tweezers (MT) to determine the effects of SYBR Gold binding on DNA length and twist. The MT assay reveals systematic lengthening and unwinding of DNA by 19.1° ± 0.7° per molecule upon binding, consistent with intercalation, similar to the related dye SYBR Green I. We complement the MT data with spectroscopic characterization of SYBR Gold. The data are well described by a global binding model for dye concentrations ≤2.5 μM, with parameters that quantitatively agree with the MT results. The fluorescence increases linearly with the number of intercalated SYBR Gold molecules up to dye concentrations of ∼2.5 μM, where quenching and inner filter effects become relevant. In summary, we provide a mechanistic understanding of DNA-SYBR Gold interactions and present practical guidelines for optimal DNA detection and quantitative DNA sensing applications using SYBR Gold.  相似文献   

10.
Detection of single DNA molecules by multicolor quantum-dot end-labeling   总被引:3,自引:0,他引:3  
Observation of DNA–protein interactions by single molecule fluorescence microscopy is usually performed by using fluorescent DNA binding agents. However, such dyes have been shown to induce cleavage of the DNA molecule and perturb its interactions with proteins. A new method for the detection of surface-attached DNA molecules by fluorescence microscopy is introduced in this paper. Biotin- and/or digoxigenin-modified DNA fragments are covalently linked at both extremities of a DNA molecule via sequence-specific hybridization and ligation. After the modified DNA molecules have been stretched on a glass surface, their ends are visualized by multicolor fluorescence microscopy using conjugated quantum dots (QD). We demonstrate that under carefully selected conditions, the position and orientation of individual DNA molecules can be inferred with good efficiency from the QD fluorescence signals alone. This is achieved by selecting QD pairs that have the distance and direction expected for the combed DNA molecules. Direct observation of single DNA molecules in the absence of DNA staining agent opens new possibilities in the fundamental study of DNA–protein interactions. This work also documents new possibilities regarding the use of QD for nucleic acid detection and analysis.  相似文献   

11.
The binding of the antibiotics netropsin and distamycin A to DNA has been studied by thermal melting, CD and sedimentation analysis. Netropsin binds strongly at antibiotic/nucleotide ratios up to at least 0.05. CD spectra obtained using DNA model polymers reveal that netropsin binds tightly to poly (dA) · poly (dT), poly (dA-dT) · poly(dA-dT) and poly (dI-dC) · poly (dI-dC) but poorly, if at all, to poly (dG) · poly (dC). Binding curves obtained with calf thymus DNA reveal one netropsin-binding site per 6.0 nucleotides (Ka=2.9 · 105 M−1); corresponding values for distamycin A are one site per 6.1 nucleotides with Ka= 11.6 · 105 M−1. Binding sites apparently involve predominantly A·T-rich sequences whose specific conformation determines their high affinity for the two antibiotics. It is suggested that the binding is stabilized primarily by hydrogen bonding and electrostatic interactions probably in the narrow groove of the DNA helix, but without intercalation. Any local structural deformation of the helix does not involve unwinding greater than approximately 3° per bound antibiotic molecule.  相似文献   

12.
DNA binding of the Type 1 DNA polymerase from Thermus aquaticus (Taq polymerase) and its Klentaq large fragment domain have been studied as a function of temperature. Equilibrium binding assays were performed from 5 to 70°C using a fluorescence anisotropy assay and from 10 to 60°C using isothermal titration calorimetry. In contrast to the usual behavior of thermophilic proteins at low temperatures, Taq and Klentaq bind DNA with high affinity at temperatures down to 5°C. The affinity is maximal at 40–50°C. The ΔH and ΔS of binding are highly temperature dependent, and the ΔCp of binding is –0.7 to –0.8 kcal/mol K, for both Taq and Klentaq, with good agreement between van’t Hoff and calorimetric values. Such a thermodynamic profile, however, is generally associated with sequence-specific DNA binding and not non- specific binding. Circular dichroism spectra show conformational rearrangements of both the DNA and the protein upon binding. The high ΔCp of Taq/Klentaq DNA binding may be correlated with structure-specific binding in analogy to sequence- specific binding, or may be a general characteristic of proteins that primarily bind non-specifically to DNA. The low temperature DNA binding of Taq/Klentaq is suggested to be a general characteristic of thermophilic DNA binding proteins.  相似文献   

13.
A fluorescence microscopy technique has been developed to visualize the behavior of individual DNA and protein molecules. Real-time direct observation of a single DNA molecule can be used to investigate the dynamics of DNA–protein interactions, such as the DNA digestion reaction by λ exonuclease. In conventional methods it is impossible to analyze the dynamics of an individual λ exonuclease molecule on a DNA because they can only observe the average behavior of a number of exonuclease molecules. Observation of a single molecule, on the other hand, can reveal processivity and binding rate of an individual exonuclease molecule. To evaluate the dynamics of λ exonuclease, a stained λ DNA molecule with one biotinylated terminal was fixed on an avidin-coated coverslip and straightened using a d.c. electric field. Microscopic observation of digestion of a straightened DNA molecule by λ exonuclease revealed that the DNA digestion rate was ~1000 bases/s and also demonstrated high processivity.  相似文献   

14.
DNA intercalators that have high affinity and slow kinetics are developed for potential DNA-targeted therapeutics. Although many natural intercalators contain multiple chiral subunits, only intercalators with a single chiral unit have been quantitatively probed. Dumbbell-shaped DNA threading intercalators represent the next order of structural complexity relative to simple intercalators, and can provide significant insights into the stereoselectivity of DNA-ligand intercalation. We investigated DNA threading intercalation by binuclear ruthenium complex [μ-dppzip(phen)4Ru2]4+ (Piz). Four Piz stereoisomers are defined by the chirality of the intercalating subunit (Ru(phen)2dppz) and the distal subunit (Ru(phen)2ip), respectively, each of which can be either right-handed (Δ) or left-handed (Λ). We used optical tweezers to measure single DNA molecule elongation due to threading intercalation, revealing force-dependent DNA intercalation rates and equilibrium dissociation constants. The force spectroscopy analysis provided the zero-force DNA binding affinity, the equilibrium DNA-ligand elongation Δxeq, and the dynamic DNA structural deformations during ligand association xon and dissociation xoff. We found that Piz stereoisomers exhibit over 20-fold differences in DNA binding affinity, from a Kd of 27 ± 3 nM for (Δ,Λ)-Piz to a Kd of 622 ± 55 nM for (Λ,Δ)-Piz. The striking affinity decrease is correlated with increasing Δxeq from 0.30 ± 0.02 to 0.48 ± 0.02 nm and xon from 0.25 ± 0.01 to 0.46 ± 0.02 nm, but limited xoff changes. Notably, the affinity and threading kinetics is 10-fold enhanced for right-handed intercalating subunits, and 2- to 5-fold enhanced for left-handed distal subunits. These findings demonstrate sterically dispersed transition pathways and robust DNA structural recognition of chiral intercalators, which are critical for optimizing DNA binding affinity and kinetics.  相似文献   

15.
Physical and linkage mapping underpin efforts to sequence and characterize the genomes of eukaryotic organisms by providing a skeleton framework for whole genome assembly. Hitherto, linkage and physical “contig” maps were generated independently prior to merging. Here, we develop a new and easy method, BAC HAPPY MAPPING (BAP mapping), that utilizes BAC library pools as a HAPPY mapping panel together with an Mbp-sized DNA panel to integrate the linkage and physical mapping efforts into one pipeline. Using Arabidopsis thaliana as an exemplar, a set of 40 Sequence Tagged Site (STS) markers spanning ∼10% of chromosome 4 were simultaneously assembled onto a BAP map compiled using both a series of BAC pools each comprising 0.7x genome coverage and dilute (0.7x genome) samples of sheared genomic DNA. The resultant BAP map overcomes the need for polymorphic loci to separate genetic loci by recombination and allows physical mapping in segments of suppressed recombination that are difficult to analyze using traditional mapping techniques. Even virtual “BAC-HAPPY-mapping” to convert BAC landing data into BAC linkage contigs is possible.  相似文献   

16.
Integrase Interactor 1 (INI1/hSNF5) is a component of the hSWI/SNF chromatin remodeling complex. The INI1 gene is either deleted or mutated in rhabdoid cancers like ATRT (Atypical terratoid and rhabdoid tumor). INI1 is also a host factor for HIV-1 replication. INI1 binds DNA non-specifically. However, the mechanism of DNA binding and its biological role are unknown. From agarose gel retardation assay (AGRA), Ni-NTA pull-down and atomic force microscopy (AFM) studies we show that amino acids 105–183 of INI1 comprise the minimal DNA binding domain (DBD). The INI1 DBD is absent in plants and in yeast SNF5. It is present in Caenorhabditis elegans SNF5, Drosophila melanogaster homologue SNR1 and is a highly conserved domain in vertebrates. The DNA binding property of this domain in SNR1, that is only 58% identical to INI1/hSNF5, is conserved. Analytical ultracentrifugation studies of INI1 DBD and INI1 DBD:DNA complexes at different concentrations show that the DBD exists as a monomer at low protein concentration and two molecules of monomer binds one molecule of DNA. At high protein concentration, it exists as a dimer and binds two DNA molecules. Furthermore, isothermal calorimetry (ITC) experiments demonstrate that the DBD monomer binds DNA with a stoichiometry (N) of ∼0.5 and Kd  = 0.94 µM whereas the DBD dimer binds two DNA molecules sequentially with K’d1 = 222 µM and K’d2 = 1.16 µM. Monomeric DBD binding to DNA is enthalpy driven (ΔH = –29.9 KJ/mole). Dimeric DBD binding to DNA is sequential with the first binding event driven by positive entropy (ΔH’1 = 115.7 KJ/mole, TΔS’1 = 136.8 KJ/mole) and the second binding event driven by negative enthalpy (ΔH’2 = –106.3 KJ/mole, TΔS’2 = –75.7 KJ/mole). Our model for INI1 DBD binding to DNA provides new insights into the mechanism of DNA binding by INI1.  相似文献   

17.
DNA bending induced by six DNA (cytosine-5) methyltransferases was studied using circular permutation gel mobility shift assay. The following bend angles were obtained: M.BspRI (GGm5CC), 46–50°; M.HaeIII (GGm5CC), 40–43°; M.SinI (GGWm5CC), 34–37°; M.Sau96I (GGNm5CC), 52–57°; M.HpaII (Cm5CGG), 30°; and M.HhaI (Gm5CGC), 13°. M.HaeIII was also tested with fragments carrying a methylated binding site, and it was found to induce a 32° bend. A phase-sensitive gel mobility shift assay, using a set of DNA fragments with a sequence-directed bend and a single methyltransferase binding site, indicated that M.HaeIII and M.BspRI bend DNA toward the minor groove. The DNA curvature induced by M.HaeIII contrasts with the lack of DNA bend observed for a covalent M.HaeIII–DNA complex in an earlier X-ray study. Our results and data from other laboratories show a correlation between the bending properties and the recognition specificities of (cytosine-5) methyltransferases: enzymes recognizing a cytosine 3′ to the target cytosine tend to induce greater bends than enzymes with guanine in this position. We suggest that the observed differences indicate different mechanisms employed by (cytosine-5) methyltransferases to stabilize the helix after the target base has flipped out.  相似文献   

18.
The Renaturation of Denatured DNA   总被引:1,自引:0,他引:1       下载免费PDF全文
The kinetics of renaturation of heat-denatured DNA from E. coli and pneumococcus have been examined by ultraviolet absorption measurements. The molecularity of the reaction was assessed by three independent treatments of the data, and all lead to the conclusion that renaturation is essentially first order at 60°; at 70° and 80° there is an increasing second order component, resulting in simultaneous unimolecular and bimolecular kinetics. The unimolecular kinetics rule out reaction between two, kinetically separate strands, indicating rather the zippering-up of a single, denatured entity. The bimolecular kinetics can be attributed to the complexing of two such entities; thus, the genetic or density-labeled complexes that have been observed by other investigators can be accounted for without invoking strand separation. Since renaturation at best is never complete, the free ends of two renatured molecules permit sufficient bimolecular reaction to produce density hybrids. The observed kinetics can be accounted for if the hydrogen bonds of DNA are broken during heat denaturation but the strands do not separate. Light scattering supports this by showing that the molecular weight is unchanged by denaturation. Since there is no existing evidence that is inconsistent with this hypothesis, it is reasonable to conclude that heat denaturation does not completely separate the entangled strands of the DNA molecule.  相似文献   

19.
Promiscuous mismatch extension by human DNA polymerase lambda   总被引:2,自引:1,他引:1  
DNA polymerase lambda (Pol λ) is one of several DNA polymerases suggested to participate in base excision repair (BER), in repair of broken DNA ends and in translesion synthesis. It has been proposed that the nature of the DNA intermediates partly determines which polymerase is used for a particular repair reaction. To test this hypothesis, here we examine the ability of human Pol λ to extend mismatched primer-termini, either on ‘open’ template-primer substrates, or on its preferred substrate, a 1 nt gapped-DNA molecule having a 5′-phosphate. Interestingly, Pol λ extended mismatches with an average efficiency of ≈10−2 relative to matched base pairs. The match and mismatch extension catalytic efficiencies obtained on gapped molecules were ≈260-fold higher than on template-primer molecules. A crystal structure of Pol λ in complex with a single-nucleotide gap containing a dG·dGMP mismatch at the primer-terminus (2.40 Å) suggests that, at least for certain mispairs, Pol λ is unable to differentiate between matched and mismatched termini during the DNA binding step, thus accounting for the relatively high efficiency of mismatch extension. This property of Pol λ suggests a potential role as a ‘mismatch extender’ during non-homologous end joining (NHEJ), and possibly during translesion synthesis.  相似文献   

20.
Ethidium bromide (EtBr) is the conventional intercalator for visualizing DNA. Previous studies suggested that EtBr lengthens and unwinds double-stranded DNA (dsDNA). However, no one has observed the unwinding of a single dsDNA molecule during intercalation. We developed a simple method to observe the twisting motions of a single dsDNA molecule under an optical microscope. A short dsDNA was attached to a glass surface of a flow chamber at one end and to a doublet bead as a rotation marker at the other end. After the addition and removal of EtBr, the bead revolved in opposite directions that corresponded to the unwinding and rewinding of a dsDNA, respectively. The amount of intercalating EtBr was estimated from the revolutions of the bead. EtBr occupied 57% of base pairs on a single dsDNA at 1 mM of EtBr, indicating that EtBr molecules could bind at contiguous sites to each other. The isotherm of intercalation showed that negative cooperativity existed between adjoining EtBr molecules. The association constant of EtBr and dsDNA (1.9 (±0.1) × 105 M−1) was consistent with that of previous results. Our system is useful to investigate the twisting of a single dsDNA interacting with various chemicals and biomolecules.  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