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 62 毫秒
1.
Four different molecular dynamics (MD) simulations have been performed for infinitely long ordered DNA molecules with different counterions, namely the two natural polyamines spermidine(3+) (Spd3+) and putrescine(2+) (Put2+), the synthetic polyamine diaminopropane(2+) (DAP2+), and the simple monovalent cation Na+. All systems comprised a periodical hexagonal cell with three identical DNA decamers, 15 water molecules per nucleotide, and counterions balancing the DNA charge. The simulation setup mimics the DNA state in oriented DNA fibers, previously studied using NMR and other experimental methods. In this paper the interplay between polyamine binding and local DNA structure is analyzed by investigating how and if the minor groove width of DNA depends on the presence and dynamics of the counterions. The results of the MD simulations reveal principal differences in the polyamine–DNA interactions between the natural [spermine(4+), Spd3+, Put2+] and the synthetic (DAP2+) polyamines.Abbreviations DAP diaminopropane - DDD Drew–Dickerson dodecamer - MD molecular dynamics - Put putrescine - RDF radial distribution function - Spd spermidine - Spm spermine  相似文献   

2.
A previous study had shown that polyamines adsorb selectively on plant cell walls according to the valence of the polyamine (Messiaen et al. 1997, Plant Physiol. 113: 387–395). In this study, the adsorption of polyamines onto isolated carrot cell walls and onto pure polygalacturonic acid was investigated in the presence of competing mono- and divalent cations (Na+ and Ca2+). Putrescine (Put2+) was unable to remove all the calcium (Ca2+) from cell walls or from polygalacturonic acid. Spermidine (Spd3+) and spermine (Spm4+) adsorbed on all galacturonates and were able to remove Ca2+ completely from both the walls and the pure polygalacturonates. Therefore, Spd3+ and Spm4+, unlike Put2+, prevented polygalacturonic acid from adopting the Ca2+-induced supramolecular conformation recognized by the 2F4 anti-pectin monoclonal antibody. We show that the signal transduction cascade otherwise initiated in plant cells by Ca2+-bound α-1,4-oligogalacturonides was indeed blocked by both Spd3+ and Spm4+, but not by Put2+. The mobilization of cytosolic free Ca2+ and the cytosolic acidification usually observed after treatment with pectic fragments did not occur and the subsequent activation of phenylalanine ammonia-lyase was suppressed. It is hypothesized that the disruption by Spd3+ and Spm4+ of the Ca2+-induced supramolecular conformation of pectic fragments was the cause of the inhibition of the pectic signal. We conclude that polyamines can act on plant cell physiology by modulating the transduction of the pectic signal. Received: 14 March 1998 / Accepted: 28 October 1998  相似文献   

3.
Biogenic polyamines are found to modulate protein synthesis at different levels. This effect may be explained by the ability of polyamines to bind and influence the secondary structure of tRNA, mRNA, and rRNA. We report the interaction between tRNA and the three biogenic polyamines putrescine, spermidine, spermine, and cobalt(III)hexamine at physiological conditions, using FTIR spectroscopy, capillary electrophoresis, and molecular modeling. The results indicated that tRNA was stabilized at low biogenic polyamine concentration, as a consequence of polyamine interaction with the backbone phosphate group. The main tRNA reactive sites for biogenic polyamine at low concentration were guanine-N7/O6, uracil-O2/O4, adenine-N3, and 2′OH of the ribose. At high polyamine concentration, the interaction involves guanine-N7/O6, adenine-N7, uracil-O2 reactive sites, and the backbone phosphate group. The participation of the polycation primary amino group, in the interaction and the presence of the hydrophobic contact, are also shown. The binding affinity of biogenic polyamine to tRNA molecule was in the order of spermine > spermidine > putrescine with KSpm = 8.7 × 105 M−1, KSpd = 6.1 × 105 M−1, and KPut = 1.0 × 105 M−1, which correlates with their positively charged amino group content. Hill analysis showed positive cooperativity for the biogenic polyamines and negative cooperativity for cobalt-hexamine. Cobalt(III)hexamine contains high- and low-affinity sites in tRNA with K1 = 3.2 × 105 M−1 and K2 = 1.7 × 105 M−1, that have been attributed to the interactions with guanine-N7 sites and the backbone PO2 group, respectively. This mechanism of tRNA binding could explain the condensation phenomenon observed at high Co(III) content, as previously shown in the Co(III)–DNA complexes.  相似文献   

4.
The effects of ions (i.e. Na+, Mg2+ and polyamines including spermidine and spermine) on the stability of various DNA oligonucleotides in solution were studied. These synthetic DNA molecules contained sequences that mimic various cellular DNA structures, such as duplexes, bulged loops, hairpins and/or mismatched base pairs. Melting temperature curves obtained from the ultraviolet spectroscopic experiments indicated that the effectiveness of the stabilization of cations on the duplex formation follows the order of spermine > spermidine > Mg2+ > Na+ > Tris–HCl buffer alone at pH 7.3. Circular dichroism spectra showed that salts and polyamines did not change the secondary structures of those DNA molecules under study. Surface plasmon resonance (SPR) observations suggested that the rates of duplex formation are independent of the kind of cations used or the structure of the duplexes. However, the rate constants of DNA duplex dissociation decrease in the same order when those cations are involved. The enhancement of the duplex stability by polyamines, especially spermine, can compensate for the instability caused by abnormal structures (e.g. bulged loops, hairpins or mismatches). The effects can be so great as to make the abnormal DNAs as stable as the perfect duplex, both kinetically and thermodynamically. Our results may suggest that the interconversion of various DNA structures can be accomplished readily in the presence of polyamine. This may be relevant in understanding the role of DNA polymorphism in cells.  相似文献   

5.
DNA and its counterions: a molecular dynamics study   总被引:9,自引:7,他引:2       下载免费PDF全文
The behaviour of mobile counterions, Na+ and K+, was analysed around a B-DNA double helix with the sequence CCATGCGCTGAC in aqueous solution during two 50 ns long molecular dynamics trajectories. The movement of both monovalent ions remains diffusive in the presence of DNA. Ions sample the complete space available during the simulation time, although individual ions sample only about one-third of the simulation box. Ions preferentially sample electronegative sites around DNA, but direct binding to DNA bases remains a rather rare event, with highest site occupancy values of <13%. The location of direct binding sites depends greatly on the nature of the counterion. While Na+ binding in both grooves is strongly sequence-dependent with the preferred binding site in the minor groove, K+ mainly visits the major groove and binds close to the centre of the oligomer. The electrostatic potential of an average DNA structure therefore cannot account for the ability of a site to bind a given cation; other factors must also play a role. An extensive analysis of the influence of counterions on DNA conformation showed no evidence of minor groove narrowing upon ion binding. A significant difference between the conformations of the double helix in the different simulations can be attributed to extensive α/γ transitions in the phosphate backbone during the simulation with Na+. These transitions, with lifetimes over tens of nanoseconds, however, appear to be correlated with ion binding to phosphates. The ion-specific conformational properties of DNA, hitherto largely overlooked, may play an important role in DNA recognition and binding.  相似文献   

6.
Binding constants for triplex formation between purine-rich oligonucleotides and a pyrimidine·purine tract of the human c-src proto-oncogene were measured by fluorescence polarization in the presence of polyamines, Na+ and K+. In both the hexamine and tetramine series, the longer polyamines had the larger binding constants for triplex formation at low concentrations of polyamine. At higher concentrations all values tended to plateau in the 109/M range. In contrast to previous reports, K+ did not inhibit triplex formation and at 150 mM the binding constants were again in the 109/M range for both an 11mer and 22mer oligonucleotide. At 150 mM K+ the addition of polyamines did not lead to any significant increase in the binding constants. It was determined that the lack of inhibition by K+ was due to the low concentration (1 nM) of purine oligonucleotide required for the fluorescence polarization technique. At higher concentrations (1 µM) self-association of the oligonucleotide was observed. These results suggest that in vivo, at least for the c-src promoter, the inhibition of triplex formation by K+ may not be detrimental. However, it may be difficult to achieve binding constants above ~109/M even in the presence of polycations.  相似文献   

7.
Four genomic DNAs of differing GC content (Micrococcus luteus, 72% GC; Escherichia coli, 50% GC; calf thymus, 42% GC; Clostridium perfringens, 27% GC) have been employed as targets of interaction by the cationic polyamines spermidine {[H3N(CH2)3NH2(CH2)4NH3]3+} and spermine {[(CH2)4(NH2(CH2)3NH3)2]4+}. In solutions containing 60 mM DNA phosphate (~20 mg DNA/ml) and either 1, 5 or 60 mM polyamine, only Raman bands associated with the phosphates exhibit large spectral changes, demonstrating that B-DNA phosphates are the primary targets of interaction. Phosphate perturbations, which are independent of base composition, are consistent with a model of non-specific cation binding in which delocalized polyamines diffuse along DNA while confined by the strong electrostatic potential gradient perpendicular to the helix axis. This finding provides experimental support for models in which polyamine-induced DNA condensation is driven by non-specific electrostatic binding. The Raman spectra also demonstrate that major groove sites (guanine N7 and thymine C5H3) are less affected than phosphates by polyamine–DNA interactions. Modest dependence of polyamine binding on genome base composition suggests that sequence context plays only a secondary role in recognition. Importantly, the results demonstrate that polyamine binding has a negligible effect on the native B-form secondary structure. The capability of spermidine or spermine to bind and condense genomic B-DNA without disrupting the native structure must be taken into account when considering DNA organization within bacterial nucleoids or cell nuclei.  相似文献   

8.
Four 20 ns molecular dynamics simulations have been performed with two counterions, K+ or Na+, at two water contents, 15 or 20 H2O per nucleotide. A hexagonal simulation cell comprised of three identical DNA decamers [d(5′-ATGCAGTCAG) × d(5′-TGACTGCATC)] with periodic boundary condition along the DNA helix was used. The simulation setup mimics the DNA state in oriented DNA fibers or in crystals of DNA oligomers. Variation of counterion nature and water content do not alter averaged DNA structure. K+ and Na+ binding to DNA are different. K+ binds to the electronegative sites of DNA bases in the major and the minor grooves, while Na+ interacts preferentially with the phosphate groups. Increase of water causes a shift of both K+ and Na+ from the first hydration shell of O1P/O2P and of the DNA bases in the minor groove with lesser influence for the cation binding to the bases in the major groove. Mobility of both water and cations in the K–DNA systems is faster than in the Na–DNA systems: Na+ organizes and immobilizes water structure around itself and near DNA while for K+ water is less organized and more dynamic.  相似文献   

9.
The effects of 3′ single-strand dangling-ends of different lengths, sequence identity of hairpin loop, and hairpin loop biotinylation at different loop residues on DNA hairpin thermodynamic stability were investigated. Hairpins contained 16 bp stem regions and five base loops formed from the sequence, 5′-TAGTCGACGTGGTCC-N5-GGACCACGTCGACTAG-En-3′. The length of the 3′ dangling-ends (En) was n = 13 or 22 bases. The identities of loop bases at positions 2 and 4 were varied. Biotinylation was varied at loop base positions 2, 3 or 4. Melting buffers contained 25 or 115 mM Na+. Average tm values for all molecules were 73.5 and 84.0°C in 25 and 115 mM Na+, respectively. Average two-state parameters evaluated from van’t Hoff analysis of the melting curve shapes in 25 mM Na+ were ΔHvH = 84.8 ± 15.5 kcal/mol, ΔSvH = 244.8 ± 45.0 cal/K·mol and ΔGvH = 11.9 ± 2.1 kcal/mol. In 115 mM Na+, two-state parameters were not very different at ΔHvH = 80.42 ± 12.74 kcal/mol, ΔSvH = 225.24 ± 35.88 cal/K·mol and ΔGvH = 13.3 ± 2.0 kcal/mol. Differential scanning calorimetry (DSC) was performed to test the validity of the two-state assumption and evaluated van’t Hoff parameters. Thermodynamic parameters from DSC measurements (within experimental error) agreed with van’t Hoff parameters, consistent with a two-state process. Overall, dangling-end DNA hairpin stabilities are not affected by dangling-end length, loop biotinylation or sequence and vary uniformly with [Na+]. Consider able freedom is afforded when designing DNA hairpins as probes in nucleic acid based detection assays, such as microarrays.  相似文献   

10.
The observation of light metal ions in nucleic acids crystals is generally a fortuitous event. Sodium ions in particular are notoriously difficult to detect because their X-ray scattering contributions are virtually identical to those of water and Na+…O distances are only slightly shorter than strong hydrogen bonds between well-ordered water molecules. We demonstrate here that replacement of Na+ by K+, Rb+ or Cs+ and precise measurements of anomalous differences in intensities provide a particularly sensitive method for detecting alkali metal ion-binding sites in nucleic acid crystals. Not only can alkali metal ions be readily located in such structures, but the presence of Rb+ or Cs+ also allows structure determination by the single wavelength anomalous diffraction technique. Besides allowing identification of high occupancy binding sites, the combination of high resolution and anomalous diffraction data established here can also pinpoint binding sites that feature only partial occupancy. Conversely, high resolution of the data alone does not necessarily allow differentiation between water and partially ordered metal ions, as demonstrated with the crystal structure of a DNA duplex determined to a resolution of 0.6 Å.  相似文献   

11.
We synthesized a pentamine (3-3-3-3) and two hexamine (3-3-3-3-3 and 3-4-3-4-3) analogues of the natural polyamine, spermine (3-4-3) and studied their effectiveness in condensing pGL3 plasmid DNA, using light scattering and atomic force microscopic (AFM) techniques. The midpoint concentration of the polyamines on pGL3 condensation (EC50) was 11.3, 10.6, 1.5, 0.49 and 0.52 µM, respectively, for 3-4-3, norspermine (3-3-3), 3-3-3-3, 3-3-3-3-3 and 3-4-3-4-3 in 10 mM Na cacodylate buffer. Dynamic laser light scattering study showed a decrease in hydrodynamic radii of plasmid DNA particles as the number of positive charges on the polyamines increased. AFM data showed the presence of toroids with outer diameter of 117–191 nm for different polyamines, and a mean height of 2.61 ± 0.77 nm. AFM results also revealed the presence of intermediate structures, including those showing circumferential winding of DNA to toroids. The dependence of the EC50 on Na+ concentration suggests different modes of binding of spermine and its higher valent analogues with DNA. Our results show a 20-fold increase in the efficacy of hexamines for DNA condensation compared to spermine, and provide new insights into the mechanism(s) of DNA nanoparticle formation. These studies might help to develop novel nonviral gene delivery vehicles.  相似文献   

12.
The interactions of self-complementary oligonucleotides with a group of metal-mediated DNA-binding drugs, including chromomycin A3, mithramycin and the novel compound UK-1, were examined via electrospray ionization quadrupole ion trap mass spectrometry. Both chromomycin and mithramycin were shown to bind preferentially to GC-rich oligonucleotide duplexes in a 2:1 drug:metal ratio, while UK-1 was shown to bind in a 1:1 drug:metal stoichiometric ratio without a strong sequence preference. These trends were observed in the presence of Co2+, Ni2+ and Zn2+, with the exception that chromomycin–Zn2+ complexes were not readily observed. The binding stoichiometries as well as the sequence specificities are in agreement with literature reports for solution studies. Binding selectivities and stabilities of the complexes were also probed using electrospray ionization mass spectrometry. Both of the GC-rich oligomers 5′-GCGCGC-3′ and 5′-GCGCATGCGC-3′ exhibited a binding preference for chromomycin over mithramycin in the presence of Co2+ and Ni2+. Energy-variable collisionally activated dissociation of the complexes was employed to determine the stabilities of the complexes. The relative metal-dependent binding energies were Ni2+ > Zn2+ > Co2+ for UK-1–oligomer complexes and Ni2+ > Co2+ for both mithramycin and chromomycin complexes.  相似文献   

13.
RecBCD is an ATP-dependent helicase and exonuclease which generates 3′ single-stranded DNA (ssDNA) ends used by RecA for homologous recombination. The exonuclease activity is altered when RecBCD encounters a Chi sequence (5′-GCTGGTGG-3′) in double-stranded DNA (ds DNA), an event critical to the generation of the 3′-ssDNA. This study tests the effect of ssDNA oligonucleotides having a Chi sequence (Chi+) or a single base change that abolishes the Chi sequence (Chio), on the enzymatic activities of RecBCD. Our results show that a 14 and a 20mer with Chi+ in the center of the molecule inhibit the exonuclease and helicase activities of RecBCD to a greater extent than the corresponding Chio oligonucleotides. Oligonucleotides with the Chi sequence at one end, or the Chi sequence alone in an 8mer, failed to show Chi-specific inhibition of RecBCD. Thus, Chi recognition requires that Chi be flanked by DNA at either end. Further experiments indicated that the oligonucleotides inhibit RecBCD from binding to its dsDNA substrate. These results suggest that a specific site for Chi recognition exists on RecBCD, which binds Chi with greater affinity than a non-Chi sequence and is probably adjacent to non-specific DNA binding sites.  相似文献   

14.
Thermodynamic measurements of ion binding to the Streptomyces lividans K+ channel were carried out using isothermal titration calorimetry, whereas atomic structures of ion-bound and ion-free conformations of the channel were characterized by x-ray crystallography. Here we use these assays to show that the ion radius dependence of selectivity stems from the channel's recognition of ion size (i.e., volume) rather than charge density. Ion size recognition is a function of the channel's ability to adopt a very specific conductive structure with larger ions (K+, Rb+, Cs+, and Ba2+) bound and not with smaller ions (Na+, Mg2+, and Ca2+). The formation of the conductive structure involves selectivity filter atoms that are in direct contact with bound ions as well as protein atoms surrounding the selectivity filter up to a distance of 15 Å from the ions. We conclude that ion selectivity in a K+ channel is a property of size-matched ion binding sites created by the protein structure.  相似文献   

15.
3′-Phosphoadenosine-5′-phosphatase (PAPase) is required for the removal of toxic 3′-phosphoadenosine-5′-phosphate (PAP) produced during sulfur assimilation in various eukaryotic organisms. This enzyme is a well-known target of lithium and sodium toxicity and has been used for the production of salt-resistant transgenic plants. In addition, PAPase has also been proposed as a target in the treatment of manic-depressive patients. One gene, halA, which could encode a protein closely related to the PAPases of yeasts and plants, was identified from the cyanobacterium Arthrospira (Spirulina) platensis. Phylogenic analysis indicated that proteins related to PAPases from several cyanobacteria were found in different clades, suggesting multiple origins of PAPases in cyanobacteria. The HalA polypeptide from A. platensis was overproduced in Escherichia coli and used for the characterization of its biochemical properties. HalA was dependent on Mg2+ for its activity and could use PAP or 3′-phosphoadenosine-5′-phosphosulfate as a substrate. HalA is sensitive to Li+ (50% inhibitory concentration [IC50] = 3.6 mM) but only slightly sensitive to Na+ (IC50 = 600 mM). The salt sensitivity of HalA was thus different from that of most of its eukaryotic counterparts, which are much more sensitive to both Li+ and Na+, but was comparable to the PAPase AtAHL (Hal2p-like protein) from Arabidopsis thaliana. The properties of HalA could help us to understand the structure-function relationship underlying the salt sensitivity of PAPases. The expression of halA improved the Li+ tolerance of E. coli, suggesting that the sulfur-assimilating pathway is a likely target of salt toxicity in bacteria as well.  相似文献   

16.
The RepA protein from bacteriophage P1 binds DNA to initiate replication. RepA covers one face of the DNA and the binding site has a completely conserved T that directly faces RepA from the minor groove at position +7. Although all four bases can be distinguished through contacts in the major groove of B-form DNA, contacts in the minor groove cannot easily distinguish between A and T bases. Therefore the 100% conservation at this position cannot be accounted for by direct contacts approaching into the minor groove of B-form DNA. RepA binding sites with modified base pairs at position +7 were used to investigate contacts with RepA. The data show that RepA contacts the N3 proton of T at position +7 and that the T=A hydrogen bonds are already broken in the DNA before RepA binds. To accommodate the N3 proton contact the T+7 /A+7 base pair must be distorted. One possibility is that T+7 is flipped out of the helix. The energetics of the contact allows RepA to distinguish between all four bases, accounting for the observed high sequence conservation. After protein binding, base pair distortion or base flipping could initiate DNA melting as the second step in DNA replication.  相似文献   

17.
Binding of monovalent and divalent cations to two adenine–adenine platform structures from the Tetrahymena group I intron ribozyme has been studied using continuum solvent models based on the generalised Born and the finite-difference Poisson–Boltzmann approaches. The adenine–adenine platform RNA motif forms an experimentally characterised monovalent ion binding site important for ribozyme folding and function. Qualitative agreement between calculated and experimental ion placements and binding selectivity was obtained. The inclusion of solvation effects turned out to be important to obtain low energy structures and ion binding placements in agreement with the experiment. The calculations indicate that differences in solvation of the isolated ions contribute to the calculated ion binding preference. However, Coulomb attraction and van der Waals interactions due to ion size differences and RNA conformational adaptation also influence the calculated ion binding affinity. The calculated alkali ion binding selectivity for both platforms followed the order K+ > Na+ > Rb+ > Cs+ > Li+ (Eisenman series VI) in the case of allowing RNA conformational relaxation during docking. With rigid RNA an Eisenman series V was obtained (K+ > Rb+ > Na+ > Cs+ > Li+). Systematic energy minimisation docking simulations starting from several hundred initial placements of potassium ions on the surface of platform containing RNA fragments identified a coordination geometry in agreement with the experiment as the lowest energy binding site. The approach could be helpful to identify putative ion binding sites in nucleic acid structures determined at low resolution or with experimental methods that do not allow identification of ion binding sites.  相似文献   

18.
The interaction of two DNA octamers, d(m5CG)4 and d(GGAATTCC), with the polyamines spermine4+ and spermidine3+, has been studied by means of 1H‐nmr nuclear Overhauser effect (NOE) difference measurements. The experiments were performed at 10°C and for a polyamine charge to DNA charge (i.e., phosphate) ratio of 0.4, where the solution of d(m5CG)4 contains about 50% Z‐form of the DNA. The results show that the polyamine intramolecular NOEs for the protons on the propyl chains are similarly negative with the two oligonucleotides, while those on the butyl chain show slightly more negative NOE with d(m5CG)4 than with d(GGAATTCC). The fully N‐methylated analogues of spermine (Me10Spn4+) and spermidine (Me8Spd3+) as well as the diamines 1,3‐diaminopropane (DAP2+) and 1,4‐diaminobutane (putrescine2+) have been studied for the ability to transform d(m5CG)4 from the B‐ to the Z‐form. 1H‐nmr spectra showed the order spermine4+ > spermidine3+ > Me10Spn4+ > Me8Spd3+ > 1,3‐diaminopropane2+ > putrescine2+, with spermine showing the largest relative amount of Z‐DNA. 1H‐nmr pulsed‐gradient self‐diffusion measurements of the triamines showed a large difference in the interaction of Spd and Me8Spd with the two different duplexes. With the same duplex (either of the two), however, no difference between Spd and Me8Spd can be seen. Within a two‐state model this is interpreted as a larger fraction of bound polyamines with d(m5CG)4 than with d(GGAATTCC). © 1999 John Wiley & Sons, Inc. Biopoly 49: 41–53, 1999  相似文献   

19.
Chloramphenicol is thought to interfere competitively with the binding of the aminoacyl-tRNA 3′-terminus to ribosomal A-site. However, noncompetitive or mixed-noncompetitive inhibition, often observed to be dependent on chloramphenicol concentration and ionic conditions, leaves some doubt about the precise mode of action. Here, we examine further the inhibition effect of chloramphenicol, using a model system derived from Escherichia coli in which a peptide bond is formed between puromycin and AcPhe-tRNA bound at the P-site of poly(U)-programmed ribosomes, under ionic conditions (6 mM Mg2+, 100 mM NH4+, 100 µM spermine) more closely resembling the physiological status. Kinetics reveal that chloramphenicol (I) reacts rapidly with AcPhe-tRNA·poly(U)·70S ribosomal complex (C) to form the encounter complex CI which is then isomerized slowly to a more tight complex, C*I. A similar inhibition pattern is observed, if complex C modified by a photoreactive analogue of spermine, reacts in buffer free of spermine. Spermine, either reversibly interacting with or covalently attached to ribosomes, enhances the peptidyltransferase activity and increases the chloramphenicol potency, without affecting the isomerization step. As indicated by photoaffinity labeling, the peptidyltransferase center at which chloramphenicol binds, is one of the preferred cross-linking sites for polyamines. This fact may explain the effect of spermine on chloramphenicol binding to ribosomes.  相似文献   

20.
Abstract

Molecular dynamics simulations with simulated annealing are performed on polyamine-DNA systems in order to determine the binding sites of putrescine, cadaverine, spermidine and spermine on A- and B-DNA. The simulations either contain no additional counterions or sufficient Na+ ions, together with the charge on the polyamine, to provide 73% neutralisation of the charges on the DNA phosphates. The stabilisation energies of the complexes indicate that all four polyamines should stabilise A-DNA in preference to B-DNA, which is in agreement with experiment in the case of spermine and spermidine, but not in the case of putrescine or cadaverine. The major groove is the preferred binding site on A-DNA of all the polyamines. Putrescine and cadaverine tend to bind to the sugar-phosphate backbone of B-DNA, whereas spermidine and spermine occupy more varied sites, including binding along the backbone and bridging both the major and minor grooves.  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