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 15 毫秒
1.
Abstract

A theoretical model is proposed for the covalent binding of (+) 7 β,8α-dihydroxy-9α, 10α- epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene denoted by BPDE I(+), to N2 on guanine. The DNA must kink a minimum of 39° to allow proper hybrid configurations about the C10 and N2 atoms involved in bond formation and to allow stacking of the pyrene moiety with the non-bonded adjacent base pair. Conservative (same sugar puckers and glycosidic angles as in B-DNA) and non-conservative (alternating sugar puckers as in intercalation sites) conformations are found and they are proposed structures in pathways connecting B-DNA, an intercalation site, and a kink site in the formation of a covalently intercalative bound adduct of BPDE I(+) to N2 on guanine. Stereographic projections are presented for (3′) and (5′) binding in the DNA. Experimental data for bending of DNA by BPDE, orientation of BPDE in DNA and unwinding of superhelical DNA is explained. The structure of a covalent intercalative complex is predicted to result from the reaction. Also, an anti ? syn transition of guanine results in a structure which allows the DNA to resume its overall B-form. The only change is that guanine has been rotated by 200° about its glycosidic bond so that the BPDE I(+) is bound in the major groove. The latter step may allow the DNA to be stored with an adduct which may produce an error in the genetic code.  相似文献   

2.
Harnessing DNA intercalation   总被引:1,自引:0,他引:1  
Numerous small molecules are known to bind to DNA through base pair intercalation. Fluorescent dyes commonly used for nucleic acid staining, such as ethidium, are familiar examples. Biological and physical studies of DNA intercalation have historically been motivated by mutation and drug discovery research. However, this same mode of binding is now being harnessed for the creation of novel molecular assemblies. Recent studies have used DNA scaffolds and intercalators to construct supramolecular assemblies that function as fluorescent 'nanotags' for cell labeling. Other studies have demonstrated how intercalators can be used to promote the formation of otherwise unstable nucleic acid assemblies. These applications illustrate how intercalators can be used to facilitate and expand DNA-based nanotechnology.  相似文献   

3.
Abstract

A nonoriented hydrated film of poly(dG-dC) with ≈20 water molecules per nucleotide (called B by Loprete and Hartman (Biochem. 32, 4077–4082 (1993)) was studied by Fourier transform infrared (FT-IR) spectroscopy either as equilibrated sample between 290 and 270 K or, after quenching into the glassy state, as nonequilibrated film isothermally at 200 and 220 K. IR spectral changes on isothermal relaxation at 200 and 220 K, caused by interconversion of two conformer substates, are revealed by difference spectra. Comparison with difference curves obtained in the same manner from two classical B-DNA forms, namely the d(CGCGAATTCGCG)2 dodecamer and polymeric NaDNA from salmon testes, revealed that the spectral changes on BIto-BII interconversion in the classical B-DNA forms are very similar to those in the B-form, and that the spectroscopic differences between the BI and BII features from classical B-DNA and those from the modified B-form are minor. Nonexponential kinetics of the BI→BII transition in the B-form of poly(dG-dC) at 200 K showed that the structural relaxation time is about three times of that in the classical B-DNA forms (≈30 versus ≈10 min at 200 K). The unexpected reversal of conformer substates interconversion (that is BII→BI transition on cooling from 290 K and BI→BII transition on isothermal relaxation at 200 K) observed for classical B-DNA occurs also in the modified B-form. We therefore conclude that restructuring of hydration shells rules the low-temperature dynamics of the B-form via its two conformer substates in the same manner reported for classical B-DNA by Pichler et al. (J. Phys. Chem. B 106, 3263–3274 (2002)).  相似文献   

4.
Daunomycin is a widely used antibiotic of the anthracycline family. In the present study we reveal the structural properties and important intercalator-DNA interactions by means of molecular dynamics. As most of the X-ray structures of DNA-daunomycin intercalated complexes are short hexamers or octamers of DNA with two drug molecules per doublehelix we calculated a self complementary 14-mer oligodeoxyribonucleotide duplex d(CGCGCGATCGCGCG)2 in the B-form with two putative intercalation sites at the 5'-CGA-3' step on both strands. Consequently we are able to look at the structure of a 1:1 complex and exclude crystal packing effects normally encountered in most of the X-ray crystallographic studies conducted so far. We performed different 10 to 20 ns long molecular dynamics simulations of the uncomplexed DNA structure, the DNA-daunomycin complex and a 1:2 complex of DNA-daunomycin where the two intercalator molecules are stacked into the two opposing 5'-CGA-3' steps. Thereby--in contrast to X-ray structures--a comparison of a complex of only one with a complex of two intercalators per doublehelix is possible. The chromophore of daunomycin is intercalated between the 5'-CG-3' bases while the daunosamine sugar moiety is placed in the minor groove. We observe a flexibility of the dihedral angle at the glycosidic bond, leading to three different positions of the ammonium group responsible for important contacts in the minor groove. Furthermore a distinct pattern of BI and BII around the intercalation site is induced and stabilized. This indicates a transfer of changes in the DNA geometry caused by intercalation to the DNA backbone.  相似文献   

5.
Some 60 years ago chemicals that intercalate between base pairs of duplex DNA were found to amplify frameshift mutagenesis. Surprisingly, the robust induction of frameshifts by intercalators still lacks a mechanistic model, leaving this classic phenomenon annoyingly intractable. A promising idea of asymmetric half‐intercalation‐stabilizing frameshift intermediates during DNA synthesis has never been developed into a model. Instead, researchers of frameshift mutagenesis embraced the powerful slipped‐mispairing concept that unexpectedly struggled with the role of intercalators in frameshifting. It is proposed that the slipped mispairing and the half‐intercalation ideas are two sides of the same coin. Further, existing findings are reviewed to test predictions of the combined “half‐intercalation into the slipped‐mispairing intermediate” model against accumulated knowledge. The existence of potential endogenous intercalators and the phenomenon of “DNA bookmarks” reveal ample possibilities for natural frameshift mutagenisis in the cell. From this alarming perspective, it is discussed how the cell could prevent genome deterioration from frameshift mutagenesis.  相似文献   

6.
Abstract

Conformational analysis has revealed anisotropic flexibility of the B-DNA double helix: it bends most easily into the grooves, being the most rigid when bent in a perpendicular direction. This result implies that DNA in a nucleosome is curved by means of relatively sharp bends (“mini-kinks”) which are directed into the major and minor grooves alternatively and separated by 5–6 base pairs. The “mini-kink” model proved to be in keeping with the x-ray structure of the B-DNA dodecamer resolved later, which exhibits two “annealed kinks”, also directed into the grooves.

The anisotropy of B DNA is sequence-dependent: the pyrimidine-purine dimers (YR) favor bending into the minor groove, and the purine-pyrimidine dinucleotides (RY), into the minor one. The RR and YY dimers appear to be the most rigid dinucleotides. Thus, a DNA fragment consisting of the interchanging oligopurine and oligopyrmidine blocks 5–6 base pairs long should manifest a spectacular curvature in solution.

Similarly, a nucleotide sequence containing the RY and YR dimers separated by a half-pitch of the double helix is the most suitable for wrapping around the nucleosomal core. Analysis of the numerous examples demonstrating the specific alignment of nucleosomes on DNA confirms this concept. So, the sequence-dependent “mechanical” properties of the double helix influence the spatial arrangement of DNA in chromatin.  相似文献   

7.
Abstract

It is now very well recognized that the DNA double helix is conformationally pluralistic and that this flexibility is derived from internal motions due to backbone torsions. But what is less apparent is that such internal motions can occur in a correlated fashion and express themselves in a wide variety of structural motifs and phenomena. For example, flexibility inherent in the DNA molecule can lead to a family of Z-DNA, LZ1 and LZ2 being the two extremes and correlated internal motion can cause LZ1?LZ2 transition. More interestingly, such motions manifest themselves as breathing modes on the DNA lattice resulting in the sequence specific intercalation sites. Following a detailed stereochemical analyses we observed that the intercalation site for ethidium is located at the dCpdG sequence of the intercalated LZ1 helix (LZ1*) while that for actinomycin D is located at the dGpdC sequence of the intercalated LZ2 helix (LZ2*). From the stereochemistry of the drug binding we make experimentally testable predictions which are in fact supported by a few recent experimental studies. These studies also show that a left-handed intercalated B-DNA model is a viable intermediate in the Z to B transition which can hold the drug with binding energy comparable to that of the intercalated right-handed B-DNA.  相似文献   

8.
9.
The binding positions and relative minimum binding energies are calculated for complexes of 9-aminoacridine, proflavine, N-methylphenanthridinium, and ethidium in theoretically determined intercalation sites in B-DNA (sites I and II) and in unconstrained dimer-duplex sites. The selection of site I in B-DNA by these compounds agrees with the theoretical interpretation of studies of unwinding angles in closed circular DNA in all cases but ethidium, which is predicted to select site II. The most stable binding positions of the acridines and ethidium in unconstrained dimer-duplex units agree with experimental results of intercalation complexes of dinucleoside monophosphate units. Base-pair specificity for Watson-Crick pairing is examined. The energy of an intercalation complex is partitioned into ΔE23, the energy required to open base pairs BP2 and BP3 in B-DNA to a site, and ΔEIn, the energy change when a free molecular intercalates. ΔE23 depends strongly on the base-pair sequence, whereas ΔEIn for the four molecules studied does not. The three most stable sequences contain (pyrimidine)p(purine) units, and this provides a rationale for the exclusive formation of crystals of intercalation complexes with these units. In spite of this selectivity, the distribution of G?C and A?T base pairs is equal for these three units and persists as the more unstable sequences are included. Therefore, specificity arises from the interaction between the base pairs and the 2′-deoxyribose 5′-monophosphate backbone for the opening of B-DNA to an intercalation site and not from the interaction between the chromophore and the DNA.  相似文献   

10.
Intercalating complexes of rhodium(III) are strong photo-oxidants that promote DNA strand cleavage or electron transfer through the double helix. The 1.2 A resolution crystal structure of a sequence-specific rhodium intercalator bound to a DNA helix provides a rationale for the sequence specificity of rhodium intercalators. It also explains how intercalation in the center of an oligonucleotide modifies DNA conformation. The rhodium complex intercalates via the major groove where specific contacts are formed with the edges of the bases at the target site. The phi ligand is deeply inserted into the DNA base pair stack. The primary conformational change of the DNA is a doubling of the rise per residue, with no change in sugar pucker from B-form DNA. Based upon the five crystallographically independent views of an intercalated DNA helix observed in this structure, the intercalator may be considered as an additional base pair with specific functional groups positioned in the major groove.  相似文献   

11.
Solid-state 2H-nmr spectra have been obtained from folded films of oriented Li- and Na-DNA molecules with the purine bases selectively deuterium labeled at the 8 position. From line shape simulations, we find that the Na-DNA sample at 75% relative humidity (rh) contains both A-DNA and surprisingly large amounts of B-DNA
  • 1 Here, B-DNA refers to “B-DNA family” (i.e. B- or C-DNA).
  • (57%). For the A-DNA component the average base tilt is 23°, and the total distribution width of tilt angles and helix axis orientations is ~ 4° (standard deviation). In the B-DNA component the base tilt is ~ 0° and the total distribution width is ~ 20°. In contrast, films of Li-DNA only exhibit the B-form line shape, consistent with a base tilt of ~ 0° and a total distribution width of base tilt angles and helix axis orientations of 9°. The nmr results that demonstrate the presence of large amounts of B-DNA in the Na-DNA sample contrast with the x-ray diffraction measurements that indicated mainly A-form. The nmr spectra are used to monitor the B-DNA content in the Na-films and to evaluate procedures for increasing the A-DNA fraction.  相似文献   

    12.
    Abstract

    Antibiotic netropsin is known to bind specifically to A and T regions in DNA; the mode of binding being non-intercalative. Obviously, H-bonding between the proton donors of netropsin and acceptors N3 of A and 02 of T comes as a strong possibility which might render this specificity. In netropsin there could be 8 proton donors: four terminal amino groups and four internal imino groups. However, methylation of the terminal amino groups does not alter the binding affinity of netropsin to DNA—but the modification of the internal imino groups significantly lowers the binding affinity. Hence, the logical conclusion is that netropsin may specifically interact with A and T through H-bonding and in order to do so, it should approach the helix from the minor groove. The present paper provides experimental data which verify the conclusion mentioned above.

    Using poly(dA-dT)? poly(dA-dT) as a model system it was observed following a thorough theoretical stereochemical analysis that netropsin could bind to -(T-A-T) sequence of the polymer in the B-form through the minor groove by forming specific B-bonding. Models could be either right or left-handed B-DNA with a mono or dinucleotide repeat.

    By monitoring the 31P signals of free poly(dA-dT) ? poly(dA-dT) and netropsin-poly(dA-dT)? poly(dA-dT) complex we show that the drug changes the DNA structure from essentially a mononucleotide repeat to that of very dominant dinucleotide repeat; however the base- pairing in the DNA-drug complex remain to be Watson-Crick. Whether H-bonding is the specific mode of interaction was judged by monitoring the imino protons of netropsin in the presence of poly(dA-dT) ? poly(dA-dT). This experiment was conducted in 90% H2O + 10% D2O Using the time-shared long pulse. It was found that exchangeable imino protons of netropsin appear in the drug-DNA complex and disappear upon increasing the D2O content; thus confirming that H-bonding is indeed the specific mode of interaction. From these and several NOE measurements, we propose a structure for poly(dA-dT)? poly(dA-dT(-netropsin complex.

    In summary, experimental data indicate that netropsin binds to poly(dA-dT)? poly(dA-dT) by forming specific hydrogen bonds and that the binding interaction causes the structure to adopt a Watson-Crick paired dinucleotide repeat motif. The proposed hydrogen bonds can form only if the drug approaches the DNA from the minor groove. Within the NMR time scale the interaction between the ligand and DNA is a fast one. From the NOE experimental data, it appears that poly(dA-dT)? poly(dA-dT) in presence of netropsin exists as an equilibrium mixture of right- and left-handed B-DNA duplexes with a dinucleotide repeat—with a predominance of the left-handed form. The last conclusion is a soft one because it was very difficult to make sure the absence of spin diffusion. In a 400 base pairs long DNA duplex- drug complex (as used in this study), equilibrium between right and left-handed helices can also mean the existence of both helical domains in the same molecule with fast interchange between these domains or/and unhindered motion/propagation of these domains along the helix axis.  相似文献   

    13.
    We have determined single crystal structures of an A-DNA decamer and a B-DNA dodecamer at 0.83 and 0.95 A, respectively. The resolution of the former is the highest reported thus far for any right-handed nucleic acid duplex and the quality of the diffraction data allowed determination of the structure with direct methods. The structures reveal unprecedented details of DNA fine structure and hydration; in particular, we have reexamined the overall hydration of A- and B-form DNA, the distribution of water around phosphate groups, and features of the water structure that may underlie the B to A transition.  相似文献   

    14.
    A stably-bound external binding site for ethidium cation in the major groove of B-form DNA is proposed. This complex is stabilized by hydrogen bonding between this ligand and the nucleophilic centers O6 and N7 of guanine, both of which are accessible via the major groove. This binding site is not the same as the well-characterized electrostatically-stabilized external binding site, but rather is seen to be a covalently bound complex which is stabilized by two hydrogen bonds between the ethidium ligand and guanine in the double stranded (ds) B-form DNA. This site [(1), R. Monaco, F. Hasheer. J Biomol Struct Dyn 10, 675 (1993)] can only exist at very low occupancy ratios. The existence of this binding site leads directly to the expectation that there will exist particular mechanistic steps along the pathway of interaction between ethidium and ds B-DNA at low and high ligand concentrations that involve this binding mode. This would not only explain observations published recently [for example, see (2-6), W. Wilson, I. Lopp. Biopolymers 18, 3025 (1979); L. Wakelin, M. Waring. J Mol Biol 144, 183-214 (1980); A. Karpetyan, N. Mehrabian, G. Terzikian, A. Antonian, P. Vardevanian, M. Frank-Kamenetshii. Proceedings of the 10th Conversation, SUNY Albany, 275 (1998); P. Vardevanyan, A. Antonyan, G. Manukyan, A. Karapetyan. Experimental and Molecular Medicine 33, 205 (2001); P. Vardevanyan, A. Antonyan, L. Minasbekan, A. Karapetyan. Proceedings of the 2002 Miami Nature Biotechnology Winter Symposium, 2(S1), 144 (2002)] but also give insight into discrepancies reported in the literature over the years by different workers studying the mechanism of interaction between ethidium and DNA. In this paper this novel binding interaction is discussed, and it is shown how the elucidation of this interaction led to the proposal of two distinct mechanisms of intercalation between ds B-DNA and ethidium cation for high and low concentrations of ligand. Modeling studies show the stability, configuration, and relative energies of this outside binding site. It is expected that this externally bound complex between ethidium cation and ds B-form DNA will be experimentally detectable using fluorescent polarization and/or linear and circular dichroism spectroscopic studies [(7, 8) E. Tuite, U. Sehlstedt, P. Hagmar, B. Norden, M. Takahashi. Euro J Biochem 243, 482-492 (1997); T. Hard. Biopolymers 26, 613-618 (1987)].  相似文献   

    15.
    Recently, the porphyrin photosensitizer meso-tetra(4-N-methyl-pyridyl)-porphine was identified as a DNA-reactive agent demonstrating both electrostatic and intercalative binding. A series of porphyrin derivatives were synthesized and studied to see if similar compounds manifested identical behavior. One derivative, meso-tetra(p-N-trimethylanilinium) porphine did not exhibit intercalation behavior but did show avid binding and novel circular dichroic features when bound to B-form DNA. At an ionic strength of 1.02, the binding constant was found to be on the order of 104 and higher at lower ionic strength. The large binding constants and induced optical activity suggest that at large porphyrin/DNA ratios the final porphyrin · DNA complex may take the form of a suprahelical helix.  相似文献   

    16.
    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   

    17.
    Abstract

    This paper describes two complexes containing N,N-dimethylproflavine and the dinucleoside monophosphate, 5-iodocytidylyl(3′-5′)guanosine (iodoCpG). The first complex is triclinic, space group PI, with unit cell dimensions a = 11.78 Å, b = 14.55 Å, c = 15.50 Å, a = 89.2°, β = 86.2°, γ = 96.4°. The second complex is monoclinic, space group P21, with a = 14.20 Å, b = 19.00 Å, c = 20.73 Å, β = 103.6°. Both structures have been solved to atomic resolution and refined by Fourier and least squares methods. The first structure has been refined anisotropically to a residual of 0.09 on 5,025 observed reflections using block diagonal least squares, while the second structure has been refined isotropically to a residual of 0.13 on 2,888 reflections with full matrix least squares. The asymmetric unit in both structures contains two dimethylproflavine molecules and two iodoCpG molecules; the first structure has 16 water molecules (a total of 134 non-hydrogen atoms), while the second structure has 18 water molecules (a total of 136 non-hydrogen atoms). Both structures demonstrate intercalation of dimethylproflavine between base-paired iodoCpG dimers. In addition, dimethylproflavine molecules stack on either side of the intercalated duplex, being related by a unit cell translation along b and a axes, respectively.

    The basic structural feature of the sugar-phosphate chains accompanying dimethylproflavine intercalation in both structures is the mixed sugar puckering pattern: C3′ endo (3′-5′) C2′ endo. This same structural information is again demonstrated in the accompanying paper, which describes a complex containing dimethylproflavine with deoxyribo-CpG.

    Similar information has already appeared for other “simple” intercalators such as ethidium, acridine orange, ellipticine, 9-aminoacridine, N-methyl-tetramethylphenanthrolinium and terpyridine platinum. “Complex” intercalators, however, such as proflavine and daunomycin, have given different structural information in model studies. We discuss the possible reasons for these differences in this paper and in the accompanying paper.  相似文献   

    18.
    19.
    Abstract

    A synthetic half-operator DNA-duplex, d(GCTACTGTATGT), containing a portion of the proposed recognition sequence (CTGT) of serveral “SOS” genes, has been synthesized. The dodecamer has been characterized through 1H-NMR spectroscopy. Complete assignment of exchangeable hydrogen bonded imino protons has been acheived by applying 1D NOE techniques and an analysis of the temperature dependence of the chemical shifts. In order to determine the specific role of the CTGT consensus sequence in the overall recognition process, the oligonucleotide duplex has been titrated with the amino terminal DNA binding domain of the LexA repressor. The observation of substantial changes of 1H-NMR chemical shifts in the imino proton region upon interaction with the protein strongly suggests that the protein binds specifically to the operator DNA. The largest deviations of 1H-NMR chemical shifts upon protein binding have been observed for protons assigned to the CTGT segment, thus strongly suggesting a direct involvement of this sequence in the binding process. At high potassium chloride concentrations the 1H-NMR chemical shift deviations are reverted which is consistent with the known drop in the affinity constant of LexA for operator DNA at high salt concentrations.  相似文献   

    20.
    Peplomycin-mediated degradation of parallel-stranded (ps) duplex was investigated. It was found that Co- and Fe-peplomycins degraded ps DNA duplex by 4'-hydrogen abstraction at 5'-GPy (pyrimidine) site in a similar manner to that of antiparallel B-DNA. While the orientation of two strands of ps and B-form DNA duplexes are reversed, peplomycin metal complex can bind to ps DNA duplex to cause oxidative DNA damage. These results indicate that peplomycin metal complex mainly interacts with one strand which is damaged.  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