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.
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   

2.
Herein we describe the first application of direct linear analysis (DLA) to the mapping of a bacterial artificial chromosome (BAC), specifically the 185.1 kb-long BAC 12M9. DLA is a single molecule mapping technology, based on microfluidic elongation and interrogation of individual DNA molecules, sequence-specifically tagged with bisPNAs. A DNA map with S/N ratio sufficiently high to detect all major binding sites was obtained using only 200 molecule traces. A new method was developed to extract an oriented map from an averaged map that included a mixture of head-first and tail-first DNA traces. In addition, we applied DLA to study the conformation and tagging of highly stretched DNA. Optimal conditions for promoting sequence-specific binding of bisPNA to an 8 bp target site were elucidated using DLA, which proved superior to electromobility shift assays. DLA was highly reproducible with a hybridized tag position localized with an accuracy of ±0.7 µm or ±2.1 kb demonstrating its utility for rapid mapping of large DNA at the single molecule level. Within this accuracy, DNA molecules, stretched to at least 85% of their contour length, were stretched uniformly, so that the map expressed in relative coordinates, was the same regardless of the molecule extension.  相似文献   

3.
Many experiments involving nucleic acids require the hybridization and ligation of multiple DNA or RNA molecules to form a compound molecule. When one of the constituents is single stranded, however, the efficiency of ligation can be very low and requires significant individually tailored optimization. Also, when the molecules involved are very long (>10 kb), the reaction efficiency typically reduces dramatically. Here, we present a simple procedure to efficiently and specifically end-join two different nucleic acids using the well-known biotin–streptavidin linkage. We introduce a two-step approach, in which we initially bind only one molecule to streptavidin (STV). The second molecule is added only after complete removal of the unbound STV. This primarily forms heterodimers and nearly completely suppresses formation of unwanted homodimers. We demonstrate that the joining efficiency is 50 ± 25% and is insensitive to molecule length (up to at least 20 kb). Furthermore, our method eliminates the requirement for specific complementary overhangs and can therefore be applied to both DNA and RNA. Demonstrated examples of the method include the efficient end-joining of DNA to single-stranded and double-stranded RNA, and the joining of two double-stranded RNA molecules. End-joining of long nucleic acids using this procedure may find applications in bionanotechnology and in single-molecule experiments.  相似文献   

4.
Although the transition from the B-DNA double helix to the A-form is essential for biological function, as shown by the existence of the A-form in many protein–DNA complexes, the dynamics of this transition has not been resolved yet. According to molecular dynamics simulations the transition is expected in the time range of a few nanoseconds. The B–A transition induced by mixing of DNA samples with ethanol in stopped flow experiments is complete within the deadtime, showing that the reaction is faster than ~0.2 ms. The reaction was resolved by an electric field jump technique with induction of the transition by a dipole stretching force driving the A- to the B-form. Poly[d(A-T)] was established as a favourable model system, because of a particularly high cooperativity of the transition and because of a spectral signature allowing separation of potential side reactions. The time constants observed in the case of poly[d(A-T)] with ~1600 bp are in the range around 10 µs. An additional process with time constants of ~100 µs is probably due to nucleation. The same time constants (within experimental accuracy ±10%) were observed for a poly[d(A-T)] sample with ~70 bp. Under low salt conditions commonly used for studies of the B–A transition, the time constants are almost independent of the ionic strength. The experimental data show that a significant activation barrier exists in the B–A transition and that the helical states are clearly separated from each other, in contrast to predictions by molecular dynamics simulations.  相似文献   

5.
A self-complementary nucleotide sequence can form both a unimolecular hairpin and a bimolecular duplex. In this study, the secondary structures of the self-complementary DNA and RNA oligonucleotides with different sequences and lengths were investigated under various solution conditions by gel electrophoresis, circular dichroism (CD) and electron paramagnetic resonance (EPR) spectroscopy and a ultraviolet (UV) melting analysis. The DNA sequences tended to adopt a hairpin conformation at low cation concentrations, but a bimolecular duplex was preferentially formed at an elevated cationic strength. On the other hand, fully matched RNA sequences adopted a bimolecular duplex regardless of the cation concentration. The thermal melting experiments indicated a greater change in the melting temperature of the bimolecular duplexes (by ~20°C) than that of the hairpin (by ~10°C) by increasing the NaCl concentration from 10 mM to 1 M. Hairpin formations were also observed for the palindrome DNA sequences derived from Escherichia coli, but association of the complementary palindrome sequences was observed when spermine, one of the major cationic molecules in a cell, existed at the physiological concentration. The results indicate the role of cations for shifting the structural equilibrium toward a nucleotide assembly and implicate nucleotide structures in cells.  相似文献   

6.
The surface properties of transfer RNA (tRNA) were analyzed using a poly(ethylene glycol)/dextran aqueous two-phase system (ATPS), where the surface net hydrophobicity (HFS) and the local hydrophobicity (LH) were evaluated based on the partition coefficient of tRNA in the ATPS. According to the evaluated HFS values, the surface of the tRNA molecule was hydrophilic at 20° -40 °C, and it became hydrophobic at 50° -80 °C because of the exposure of the intrinsic nucleobases of tRNA. In contrast, the LH values were found to be maximal at 20° -40 °C. The conformation of tRNA was investigated by Raman and circular dichroism (CD) spectroscopies, corroborating the results with the calculated prediction of its secondary structure (Mfold). It was shown that 66% of A-form structure existed at room temperature; the base stacking (θ265) was gradually decreased, and the A-form structure (θ208) was denatured along with a sigmoid curve against the temperature increase; the denatured secondary structures were observed above 50° C by Mfold prediction. The HFS value of the DNA duplex was found to be hydrophilic, compared to that of the single-stranded DNA, indicating that the exposure of nucleobases is a key factor of the hydrophobic properties of nucleotides. We conclude that the hydrophobic property of the tRNA surface was directly affected by its conformational transition.  相似文献   

7.
The crystal structure of the DNA dodecamer duplex CATGGGCCCATG lies on a structural continuum along the transition between A- and B-DNA. The dodecamer possesses the normal vector plot and inclination values typical of B-DNA, but has the crystal packing, helical twist, groove width, sugar pucker, slide and x-displacement values typical of A-DNA. The structure shows highly ordered water structures, such as a double spine of water molecules against each side of the major groove, stabilizing the GC base pairs in an A-like conformation. The different hydration of GC and AT base pairs provides a physical basis for solvent-dependent facilitation of the A↔B helix transition by GC base pairs. Crystal structures of CATGGGCCCATG and other A/B-DNA intermediates support a ‘slide first, roll later’ mechanism for the B→A helix transition. In the distribution of helical parameters in protein–DNA crystal structures, GpG base steps show A-like properties, reflecting their innate predisposition for the A conformation.  相似文献   

8.
Four-way junctions (4Hs) are important intermediates in DNA rearrangements such as genetic recombination. Under the influence of multivalent cations these molecules undergo a conformational change, from an extended planar form to a quasi-continuous stacked X-structure. Recently, a number of X-ray structures and a nuclear magnetic resonance (NMR) structure of 4Hs have been reported and in three of these the position of multivalent cations is revealed. These structures belong to two main families, characterized by the angle between the two co-axial stacked helices, which is either around +40 to +55° or around –70 to –80°. To investigate the role of metal-ion binding on the conformation of folded 4Hs we performed Brownian-dynamics simulations on the set of available structures. The simulations confirm the proposed metal-ion binding sites in the NMR structure and in one of the X-ray structures. Furthermore, the calculations suggest positions for metal-ion binding in the other X-ray structures. The results show a striking dependence of the ion density on the helical environment (B-helix or A-helix) and the structural family.  相似文献   

9.
Understanding the function of complex RNA molecules depends critically on understanding their structure. However, creating three-dimensional (3D) structural models of RNA remains a significant challenge. We present a protocol (the nucleic acid simulation tool [NAST]) for RNA modeling that uses an RNA-specific knowledge-based potential in a coarse-grained molecular dynamics engine to generate plausible 3D structures. We demonstrate NAST's capabilities by using only secondary structure and tertiary contact predictions to generate, cluster, and rank structures. Representative structures in the best ranking clusters averaged 8.0 ± 0.3 Å and 16.3 ± 1.0 Å RMSD for the yeast phenylalanine tRNA and the P4-P6 domain of the Tetrahymena thermophila group I intron, respectively. The coarse-grained resolution allows us to model large molecules such as the 158-residue P4-P6 or the 388-residue T. thermophila group I intron. One advantage of NAST is the ability to rank clusters of structurally similar decoys based on their compatibility with experimental data. We successfully used ideal small-angle X-ray scattering data and both ideal and experimental solvent accessibility data to select the best cluster of structures for both tRNA and P4-P6. Finally, we used NAST to build in missing loops in the crystal structures of the Azoarcus and Twort ribozymes, and to incorporate crystallographic data into the Michel–Westhof model of the T. thermophila group I intron, creating an integrated model of the entire molecule. Our software package is freely available at https://simtk.org/home/nast.  相似文献   

10.
The solution structure and hydration of the chimeric duplex [d(CGC)r(aaa)d(TTTGCG)]2, in which the central hybrid segment is flanked by DNA duplexes at both ends, was determined using two-dimensional NMR, simulated annealing and restrained molecular dynamics. The solution structure of this chimeric duplex differs from the previously determined X-ray structure of the analogous B-DNA duplex [d(CGCAAATTTGCG)]2 as well as NMR structure of the analogous A-RNA duplex [r(cgcaaauuugcg)]2. Long-lived water molecules with correlation time τc longer than 0.3 ns were found close to the RNA adenine H2 and H1′ protons in the hybrid segment. A possible long-lived water molecule was also detected close to the methyl group of 7T in the RNA–DNA junction but not with the other two thymines (8T and 9T). This result correlates with the structural studies that only DNA residue 7T in the RNA–DNA junction adopts an O4′-endo sugar conformation, while the other DNA residues including 3C in the DNA–RNA junction, adopt C1′-exo or C2′-endo conformations. The exchange rates for RNA C2′-OH were found to be ~520 s–1. This slow exchange rate may be due to the narrow minor groove width of [d(CGC)r(aaa)d(TTTGCG)]2, which may trap the water molecules and restrict the dynamic motion of hydroxyl protons. The minor groove width of [d(CGC)r(aaa)d(TTTGCG)]2 is wider than its B-DNA analog but narrower than that of the A-RNA analog. It was further confirmed by its titration with the minor groove binding drug distamycin. A possible 2:1 binding mode was found by the titration experiments, suggesting that this chimeric duplex contains a wider minor groove than its B-DNA analog but still narrow enough to hold two distamycin molecules. These distinct structural features and hydration patterns of this chimeric duplex provide a molecular basis for further understanding the structure and recognition of DNA·RNA hybrid and chimeric duplexes.  相似文献   

11.
Sulforaphane (SFN) is an isothiocyanate found in cruciferous vegetables with anti-inflammatory, anti-oxidant and anti-cancer activities. However, the antioxidant and anticancer mechanism of sulforaphane is not well understood. In the present research, we reported binding modes, binding constants and stability of SFN–DNA and -RNA complexes by Fourier transform infrared (FTIR) and UV–Visible spectroscopic methods. Spectroscopic evidence showed DNA intercalation with some degree of groove binding. SFN binds minor and major grooves of DNA and backbone phosphate (PO2), while RNA binding is through G, U, A bases with some degree of SFN–phosphate (PO2) interaction. Overall binding constants were estimated to be K(SFN–DNA)=3.01 (± 0.035)×104 M-1 and K(SFN–RNA)= 6.63 (±0.042)×103 M-1. At high SFN concentration (SFN/RNA = 1/1), DNA conformation changed from B to A occurred, while RNA remained in A-family structure.  相似文献   

12.
Base extrusion is a major structural feature at the junction between B- and Z-DNA (the B–Z junction) where a base pair is broken, and the two bases are extruded from the double helix. Despite the demonstration of base extrusion at the B–Z junction, it is not clear whether a similar base extrusion occurs at other types of junctions involving the left-handed Z conformation. Here, we investigate structural changes of bases at three Z-form junctions: DNA B–Z and Z–Z and RNA A–Z junctions. By monitoring fluorescently labeled duplex nucleic acids using 2-aminopurines at various positions relative to the junction point, we show that base extrusion occurs not only at the DNA B–Z junction, but also at the RNA A–Z and DNA Z–Z junctions. Our data suggest that base extrusion is a general feature of Z-form nucleic-acid junctions.  相似文献   

13.
The cissyn cyclobutane pyrimidine dimer (CPD) is a cytotoxic, mutagenic and carcinogenic DNA photoproduct and is repaired by the nucleotide excision repair (NER) pathway in mammalian cells. The XPC–hHR23B complex as the initiator of global genomic NER binds to sites of certain kinds of DNA damage. Although CPDs are rarely recognized by the XPC–hHR23B complex, the presence of mismatched bases opposite a CPD significantly increased the binding affinity of the XPC–hHR23B complex to the CPD. In order to decipher the properties of the DNA structures that determine the binding affinity for XPC–hHR23B to DNA, we carried out structural analyses of the various types of CPDs by NMR spectroscopy. The DNA duplex which contains a single 3′ T·G wobble pair in a CPD (CPD/GA duplex) induces little conformational distortion. However, severe distortion of the helical conformation occurs when a CPD contains double T·G wobble pairs (CPD/GG duplex) even though the T residues of the CPD form stable hydrogen bonds with the opposite G residues. The helical bending angle of the CPD/GG duplex was larger than those of the CPD/GA duplex and properly matched CPD/AA duplex. The fluctuation of the backbone conformation and significant changes in the widths of the major and minor grooves at the double T·G wobble paired site were also observed in the CPD/GG duplex. These structural features were also found in a duplex that contains the (6–4) adduct, which is efficiently recognized by the XPC–hHR23B complex. Thus, we suggest that the unique structural features of the DNA double helix (that is, helical bending, flexible backbone conformation, and significant changes of the major and/or minor grooves) might be important factors in determining the binding affinity of the XPC–hHR23B complex to DNA.  相似文献   

14.
Jain SS  Polak M  Hud NV 《Nucleic acids research》2003,31(15):4608-4615
Small molecules that intercalate in DNA and RNA are powerful agents for controlling nucleic acid structural transitions. We recently demonstrated that coralyne, a small crescent-shaped molecule, can cause the complete and irreversible disproportionation of duplex poly(dA)·poly(dT) into triplex poly(dA)·poly(dT)·poly(dT) and a poly(dA) self- structure. Both DNA secondary structures that result from duplex disproportionation are stabilized by coralyne intercalation. In the present study, we show that the kinetics and thermodynamics of coralyne-driven duplex disproportionation strongly depend on oligonucleotide length. For example, disproportionation of duplex (dA)16·(dT)16 by coralyne reverts over the course of hours if the sample is maintained at 4°C. Coralyne-disproportioned (dA)32· (dT)32, on the other hand, only partially reverts to the duplex state over the course of days at the same temperature. Furthermore, the equilibrium state of a (dA)16·(dT)16 sample in the presence of coralyne at room temperature contains three different secondary structures [i.e. duplex, triplex and the (dA)16 self-structure]. Even the well-studied process of triplex stabilization by coralyne binding is found to be a length-dependent phenomenon and more complicated than previously appreciated. Together these observations indicate that at least one secondary structure in our nucleic acid system [i.e. duplex, triplex or (dA)n self-structure] binds coralyne in a length-dependent manner.  相似文献   

15.
NMR solution structure of a parallel LNA quadruplex   总被引:3,自引:2,他引:1  
The solution structure of a locked nucleic acid (LNA) quadruplex, formed by the oligomer d(TGGGT), containing only conformationally restricted LNA residues is reported. NMR and CD spectroscopy, as well as molecular dynamics and mechanic calculations, has been used to characterize the complex. The molecule adopts a parallel stranded conformation with a 4-fold rotational symmetry, showing a right-handed helicity and the guanine residues in an almost planar conformation with three well-defined G-tetrads. The thermal stability of Q-LNA has been found to be comparable with that of [r(UGGGU)]4, while a Tm increment of 20°C with respect to the corresponding DNA quadruplex structure [d(TGGGT)]4 has been observed. The structural features of the LNA quadruplex reported here may open new perspectives for the biological application of LNAs as novel versatile tools to design aptamer or catalyst oligonucleotides.  相似文献   

16.
RNA molecules play different roles in coding, decoding and gene expression regulation. Such roles are often associated to the RNA secondary or tertiary structures. The folding dynamics lead to multiple secondary structures of long RNA molecules, since an RNA molecule might fold into multiple distinct native states. Despite an ensemble of different structures, it has been theoretically proposed that the separation between the 5′ and 3′ ends of long single-stranded RNA molecules (ssRNA) remains constant, independent of their base content and length. Here, we present the first experimental measurements of the end-to-end separation in long ssRNA molecules. To determine this separation, we use single molecule Fluorescence Resonance Energy Transfer of fluorescently end-labeled ssRNA molecules ranging from 500 to 5500 nucleotides in length, obtained from two viruses and a fungus. We found that the end-to-end separation is indeed short, within 5–9 nm. It is remarkable that the separation of the ends of all RNA molecules studied remains small and similar, despite the origin, length and differences in their secondary structure. This implies that the ssRNA molecules are ‘effectively circularized’ something that might be a general feature of RNAs, and could result in fine-tuning for translation and gene expression regulation.  相似文献   

17.
Urea PAGE or denaturing urea polyacrylamide gel electrophoresis employs 6-8 M urea, which denatures secondary DNA or RNA structures and is used for their separation in a polyacrylamide gel matrix based on the molecular weight. Fragments between 2 to 500 bases, with length differences as small as a single nucleotide, can be separated using this method1. The migration of the sample is dependent on the chosen acrylamide concentration. A higher percentage of polyacrylamide resolves lower molecular weight fragments. The combination of urea and temperatures of 45-55 °C during the gel run allows for the separation of unstructured DNA or RNA molecules.In general this method is required to analyze or purify single stranded DNA or RNA fragments, such as synthesized or labeled oligonucleotides or products from enzymatic cleavage reactions.In this video article we show how to prepare and run the denaturing urea polyacrylamide gels. Technical tips are included, in addition to the original protocol 1,2.  相似文献   

18.
Despite considerable interest and investigations on cationic lipid–DNA complexes, reports on lipid–RNA interaction are very limited. In contrast to lipid–DNA complexes where lipid binding induces partial B to A and B to C conformational changes, lipid–tRNA complexation preserves tRNA folded state. This study is the first attempt to investigate the binding of cationic lipid with transfer RNA and the effect of lipid complexation on tRNA aggregation and condensation. We examine the interaction of tRNA with cholesterol (Chol), 1,2-dioleoyl-3-trimethylammonium-propane (DOTAP), dioctadecyldimethylammoniumbromide (DDAB) and dioleoylphosphatidylethanolamine (DOPE), at physiological condition, using constant tRNA concentration and various lipid contents. FTIR, UV-visible, CD spectroscopic methods and atomic force microscopy (AFM) were used to analyze lipid binding site, the binding constant and the effects of lipid interaction on tRNA stability, conformation and condensation. Structural analysis showed lipid–tRNA interactions with G–C and A–U base pairs as well as the backbone phosphate group with overall binding constants of KChol = 5.94 (± 0.8) × 104 M–1, KDDAB = 8.33 (± 0.90) × 105 M–1, KDOTAP = 1.05 (± 0.30) × 105 M–1 and KDOPE = 2.75 (± 0.50) × 104 M–1. The order of stability of lipid–tRNA complexation is DDAB > DOTAP > Chol > DOPE. Hydrophobic interactions between lipid aliphatic tails and tRNA were observed. RNA remains in A-family structure, while biopolymer aggregation and condensation occurred at high lipid concentrations.  相似文献   

19.
20.
In 1967 Wang and Schwartz reported on the formation of interlocked rings between linear or circular DNA molecules by the enzyme topoisomerase. We propose viewing the secondary structured loop in RNA (or single stranded DNA) as analogous to a circular DNA molecule. Formation of a catenane between such an RNA loop with a DNA molecule may constitute a probe of the secondary and general three dimensioanl structure of the RNA molecule. The experimental results may be compared with the theoretical calculation. We suggest here a method for estimating linkage probabilities and calculate them for several cases for which secondary structures of the RNA have been proposed.  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